23.08.2013 Views

pdf, 57.71Mb - Entomological Society of Canada

pdf, 57.71Mb - Entomological Society of Canada

pdf, 57.71Mb - Entomological Society of Canada

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

Biological Control<br />

Programmes against<br />

Insects and Weeds<br />

in <strong>Canada</strong> 1969-1980<br />

EDITED BY J.S. KELLEHER AND M.A. HULME<br />

@ COMMONWEALTH AGRICULTURAL BUREAUX


Commonwealth Agricultural Bureaux<br />

Farnham Royal<br />

Slough SL2 3BN<br />

England<br />

Tel. Farnham Common 2281<br />

Telex 847964<br />

ISBN 085198536 X<br />

© Commonwealth Agricultural Bureaux, 1984. All rights<br />

reserved. No part <strong>of</strong> this publication may be reproduced in any<br />

form or by any means, electronically, mechanically, by photocopying,<br />

recording or otherwise, without the prior permission<br />

<strong>of</strong> the copyright owner.<br />

The Executive Council <strong>of</strong> the Commonwealth Agricultural<br />

Bureaux is signatory to the Fair Copying Declaration, details<br />

<strong>of</strong> which can be obtained from The Royal <strong>Society</strong>, 6 Carlton<br />

House Terrace, London SWI Y SAG.<br />

Printed by Page Bros (Norwich) Ltd


CONTENTS<br />

Contents iii<br />

Page<br />

FOREWORD vii<br />

GENERAL INTRODUCTION ix<br />

PART I - Biological control <strong>of</strong> agricultural insects in <strong>Canada</strong>, 1969-1980<br />

Chapter 1. Current approaches to biological control <strong>of</strong> agricultural insect pests 3<br />

Chapter 2. Acyrthosiphon pisum (Harris), Pea Aphid (Homoptera: Aphididae) 7<br />

Chapter 3. Adelphocoris lineolatus (Goeze), Alfalfa Plant Bug (Heteroptera: Miridae) 9<br />

Chapter 4. Agromyza frontella (Rondani), Alfalfa Blotch Leafminer (Diptera: 11<br />

Agromyzidae)<br />

Chapter 5. Artogeia rapae (L.), Imported Cabbageworm (Lepidoptera: Pieridae), 15<br />

Trichoplusia ni (HObner), Cabbage Looper (Lepidoptera:Noctuidae),<br />

and Plutella xylostella (L.), Diamondback Moth (Lepidoptera:<br />

Plutellidae)<br />

Chapter 6. Culex pipiens L., Northern House Mosquito (Diptera: Culicidae) 19<br />

Chapter 7. Culiseta inornata (Williston) a Mosquito (Diptera: Culicidae) 23<br />

Chapter 8. Cydia pomonella (L.), Codling Moth (Lepidoptera: Tortricidae) 25<br />

Chapter 9. Delia antiqua (Meigen), Onion Maggot (Diptera: Anthomyiidae) 29<br />

Chapter 10. Entomoscelis americana Brown, Red Turnip Beetle (Coleoptera: 31<br />

Chrysomelidae)<br />

Chapter 11. Euxoa messoria (Harris), Darksided Cutworm (Lepidoptera: Noctuidae) 33<br />

Chapter 12. Forficula auricularia L., European Earwig (Dermaptera: Forficulidae) 39<br />

Chapter 13. Hypera postica (Gyllenhal), Alfalfa Weevil (Coleoptera: Curculionidae) 41<br />

Chapter 14. Lygus spp., Plant Bugs (Heteroptera: Miridae) 45<br />

Chapter 15. Mamcstra configurata Walker, Bertha Armyworm (Lepidoptera: Noctuidae) 49<br />

Chapter 16. Manduca quinquemacu/ata (Haworth), Tomato Hornworm (Lepidoptera: 57<br />

Sphingidae)<br />

Chapter 17. Melanoplus spp., Camnu/a pellucids (Scudder), and other Grasshoppers 61<br />

(Orthoptera: Acrididae)<br />

Chapter 18. Musca domestics L., House Fly (Diptera: Muscidae) 63<br />

Chapter 19. Dulema melanopus (L.), Cereal Leaf Beetle (Coleoptera: Chrysomelidae) 65<br />

Chapter 20. Phyllonorycter blancardella (F.), Spotted Tentiform Leafminer, (Lepidoptera: 69<br />

Gracillariidae)<br />

Chapter 21. Phyllotreta spp., Flea Beetles (Coleoptera: Chrysomelidae) 73<br />

Chapter 22. Tetranychus urticae Koch, Twospotted Spider mite (Acarina: Tetranychidae) 77<br />

Chapter 23. Thymelicus lineola (Ochsenheimer), European Skipper (Lepidoptera: 79<br />

Hesperiidae)<br />

Chapter 24. Tipu/a pa/udosa Meigen, European Cranefly (Diptera: Tipulidae) 85<br />

Chapter 25. Tria/eurodes vaporariorum (Westwood), Greenhouse Whitefly (Homoptera: 89<br />

Aleyrodidae)


iv Contents<br />

PART II - Biological control <strong>of</strong> weeds in <strong>Canada</strong>, 1969-1980<br />

Chapter 26. Current Approaches to Biological Control <strong>of</strong> Weeds<br />

Chapter 27. Acroptilon repens (L.) DC., Russian Knapweed (Compositae)<br />

Chapter 28. Ambrosia artemisiilolia L., Common Ragweed (Compositae)<br />

Chapter 29. Artemisia absinthium L., Absinth (Compositae)<br />

Chapter 30. Carduus nutans L., Nodding Thistle and C. acanthoides L., Plumeless<br />

Thistle (Compositae)<br />

Chapter 31. Centaurea diffusa Lam. and C. maculosa Lam., Diffuse and Spotted<br />

Knapweed (Compositae)<br />

Chapter 32. Cirsium arvense (L.) Scop., <strong>Canada</strong> Thistle (Compositae)<br />

Chapter 33. Clrsium vulgare (Savi) Ten., Bull Thistle (Compositae)<br />

Chapter 34. Convolvulus arvensis L., Field Bindweed (Convolvulaceae)<br />

Chapter 35. Euphorbia esu/a-virgata complex, Leafy Spurge and E. cyparissias L.,<br />

Cypress Spurge (Euphorbiaceae)<br />

Chapter 36. Hypericum perforatum L., St. John's-wort (Hypericaceae)<br />

Chapter 37. Linaria vulgaris Miller, Yellow Toadflax, and L. dalmatica (L.) Mill.,<br />

Broad-leaved ToadOax (Scrophulariaceae)<br />

Chapter 38. Opuntia polyacantha Haworth, Plains Prickly-pear Cactus (Cactaceae)<br />

Chapter 39. Rhamnus cathartica L., Common or European Buckthorn (Rhamnaceae)<br />

Chapter 40. Salso/a pest/fer A. Nels., Russian Thistle (Chenopodiaceae)<br />

Chapter 41. Senecio jacobaea L., Tansy Ragwort (Compositae)<br />

Chapter 42. Silene cucubalus Wibel, Bladder Campion (Caryophyllaceae)<br />

Chapter 43. Sonchus arvensis L., Perennial Sow-thistle, S. oleraceus L., Annual Sowthistle,<br />

S. asper (L.) Hill, Spiny Annual Sow-thistle (Compositae)<br />

Chapter 44. Verbascum thapsus L., Common Mullein (Scrophulariaceae)<br />

PART III - Biological control <strong>of</strong> forest insect pests in <strong>Canada</strong> 1969-80<br />

Chapter 45. Biological Control <strong>of</strong> Forest Insect Pests in <strong>Canada</strong> 1969-1980: 215<br />

Retrospect and Prospect<br />

Chapter 46. Adelges piceae (Ratz.), Balsam Woolly Adelgid (Homoptera: Adelgidae) 229<br />

Chapter 47. Choristoneura fumiferana (Clemens), Spruce Budworm (Lepidoptera: 235<br />

Tortricidae)<br />

A. Field Development <strong>of</strong> Bacillus thuringiensis Berliner in Eastern <strong>Canada</strong>, 238<br />

1970-80<br />

B. Viruses: Application and Assessment 248<br />

C. Application <strong>of</strong> Microsporidia and Fungi, and <strong>of</strong> Genetic Manipulation 260<br />

D. Testing <strong>of</strong> Parasitoids 267<br />

E. Review <strong>of</strong> Biological Control Opportunities against Spruce Budworm, 273<br />

Choristoneura fumiferana<br />

Chapter 48. Choristoneura occidentalls Freeman, Western Spruce Budworm (Lepidoptera: 277<br />

Tortricidae)<br />

Chapter 49. Coleophora laricella (Hubner), Larch Casebearer (Lepidoptera: 281<br />

Coleophoridae)<br />

95<br />

105<br />

111<br />

113<br />

115<br />

127<br />

139<br />

147<br />

155<br />

159<br />

171<br />

179<br />

183<br />

185<br />

191<br />

195<br />

203<br />

205<br />

211


Chapter 50. Coleophora serratella (L.), Birch Casebearer (Lepidoptera: Coleophoridae)<br />

Chapter 51. Fenusa pusilla (Lepeletier), Birch Leafminer (Hymenoptera:<br />

Tenthredinidae)<br />

Chapter 52. Gilpinia hercyniae (Hartig), European Spruce Sawfly (Hymenoptera:<br />

Diprionidae)<br />

Chapter 53. Leucoma sa/icis (L.), Satin Moth (Lepidoptera: Lymantriidae)<br />

Chapter 54. Lymantria dispar (L.), Gypsy Moth (Lepidoptera: Lymantriidae)<br />

Chapter 55. Malacosoma disstria HObner, Forest Tent Caterpillar (Lepidoptera:<br />

Lasiocampidae)<br />

Chapter 56. Neodiprion abietis (Harris), Balsam Fir Sawfly (Hymenoptera: Diprionidae)<br />

Chapter 57. Neodiprion lecontei (Fitch), Redheaded Pine Sawfly (Hymenoptera:<br />

Diprionidae)<br />

Chapter 58. Neodiprion sertifer (Ge<strong>of</strong>froy), European Pine Sawfly (Hymenoptera:<br />

Diprionidae)<br />

Chapter 59. Neodiprion swainei (Middleton), Swaine Jack Pine Sawfly (Hymenoptera:<br />

Diprionidae)<br />

Chapter 60. Operophtera bruceata (Hulst), Bruce Span worm (Lepidoptera: Geometridae)<br />

Chapter 61. Operophtera brumata (L.), Winter Moth (Lepidoptera: Geometridae)<br />

Chapter 62. Orgyia leucostigma (J.E. Smith), Whitemarked Tussock Moth (Lepidoptera:<br />

Lymantriidae)<br />

Chapter 63. Orgyia pseudotsugata (McDunnough), Douglas-fir Tussock Moth<br />

(Lepidoptera: Lymantriidae)<br />

Chapter 64. Pristiphora erichsonii (Hartig), Larch Sawfly (Hymenoptera: Tenthredinidae)<br />

Chapter 65. Pristiphora geniculata (Hartig), Mountain-ash Sawfly (Hymenoptera:<br />

Tenthredinidae)<br />

Chapter 66. Rhyacionia buoliana (Schiff.), European Pine Shoot Moth (Lepidoptera:<br />

Tortricidae)<br />

APPENDIX - Addresses <strong>of</strong> contributors<br />

INDEX OF SPECIES<br />

Contents v<br />

285<br />

291<br />

295<br />

299<br />

303<br />

311<br />

321<br />

323<br />

331<br />

341<br />

349<br />

353<br />

359<br />

363<br />

369<br />

381<br />

387<br />

395<br />

397


Foreword<br />

Foreword vii<br />

The CIBC Technical Communication series was initiated to provide an outlet for<br />

regional reviews on biological control. This review, the 8th in the series and the 3rd prepared<br />

by Canadian scientists, points to the importance which <strong>Canada</strong> attaches to biological<br />

control and reflects the volume <strong>of</strong> Canadian activity in this field.<br />

<strong>Canada</strong> continues to support an active classical biological control programme, relying<br />

on the Commonwealth Institute <strong>of</strong> Biological Control to undertake most <strong>of</strong> its foreign<br />

exploration and research. The collection and redistribution <strong>of</strong> natural enemies successfully<br />

established earlier to hasten control has continued. However, the current volume<br />

indicates an acceptance <strong>of</strong> biological control in its wider context as a component,<br />

albeit a very important one, <strong>of</strong> integrated pest management. Increasing attention and<br />

interest are being focused on other biological control approaches, i.e., augmentative<br />

and inundative releases as well as amelioration <strong>of</strong> the environment. The value <strong>of</strong> food<br />

plants or nectar sources for adult parasitoids has been demonstrated.<br />

To date, augmentative and inundative releases in <strong>Canada</strong> have involved pathogens,<br />

both introduced and native, rather than parasitoids or predators. The period under<br />

review has seen the implementation, by large field trials, <strong>of</strong> microbial control, utilizing<br />

both bacterial and viral preparations for the control <strong>of</strong> forestry pests. The use <strong>of</strong><br />

Bacillus thuringiensis has involved the large-scale production <strong>of</strong> "tailor-made" formulations<br />

by the agro-chemical industry and provides a preferred, although more<br />

expensive, alternative to the former wide-scale use <strong>of</strong> chemical pesticides for the<br />

control <strong>of</strong> spruce budworms. B.thuringiensis is fully registered as a pesticide in<br />

<strong>Canada</strong> for use against several lepidopteran defoliators. Similarly, research on viruses<br />

<strong>of</strong> neodiprionid sawflies has led to practical field application, with some degree <strong>of</strong><br />

carry-over from year to year.<br />

The first book reported introductions <strong>of</strong> control agents against two weeds; the<br />

second reviewed attempts against four additional species or species complexes<br />

whereas in the present volume, thirty agents have now been released against fourteen<br />

weed species in <strong>Canada</strong>. Increasing efforts have been made to evaluate or quantify the<br />

gains which have been or might be derived from biological control <strong>of</strong> weeds.<br />

The book is timely; by its preparation, scientists and administrators can take stock <strong>of</strong><br />

achievements to date and thus will be better placed to plan for the future. As M.A.<br />

Hulme and G.A. Green point out, the scope for research opportunities is limited not by<br />

lack <strong>of</strong> suitable problems but by the resources available to devote to biological control.<br />

The economic climate in <strong>Canada</strong>, as elsewhere, dictates that emphasis be given to<br />

developing management strategies for the pests causing the greatest economic losses<br />

and, as pointed out by Dr. P. Harris, proposals for future funding for biological control<br />

must, in order to convince administrators to release funds, provide quantitive<br />

assessment <strong>of</strong> the losses inflicted by the target pest or weed and thereby demonstrate<br />

the financial benefit which will result if the project is successful. This book attempts to<br />

provide an evaluation <strong>of</strong> each control attempt, and administrators and scientists now<br />

should be in a good position to plan for the future.<br />

F.D. Bennett<br />

Director<br />

Commonwealth Institute <strong>of</strong><br />

Biological Control


General Introduction<br />

Gcncrallntroduction ix<br />

This book is the third review <strong>of</strong> Canadian biological control work published by the<br />

Commonwealth Agricultural Bureaux. It is compiled and edited by Dr. M.A. Hulme<br />

and Dr. J .S. Kelleher. Dr. P. Harris coordinated preparation <strong>of</strong> the section on biological<br />

control <strong>of</strong> weeds.<br />

The first compilation, "A Review <strong>of</strong> the Biological Control Attempts Against Insects<br />

and Weeds in <strong>Canada</strong>" was published in 1962 and dealt with control attempts against<br />

agricultural pests and weeds up to 1959 and against forest insects from 1910-1958. In<br />

1971 a sequel was published entitled ''Biological Control Programmes against Insects<br />

and Weeds in <strong>Canada</strong> 1959-1968"; this volume had separate sections on agricultural<br />

insects, on weeds, on forest pests, and on the organization <strong>of</strong> Canadian biological<br />

control research. During the late 1970s, a further sequel was proposed and preparation<br />

began in 1980. This volume is the end result.<br />

Like its immediate predecessor, this review has two main functions. First, it collates<br />

Canadian biological control attempts against insects and weeds from 1969 to 1980, and<br />

thus provides a permanent record <strong>of</strong> otherwise diffuse and relatively inaccessible data.<br />

Secondly, it attempts to provide an evaluation <strong>of</strong> each control attempt in terms <strong>of</strong><br />

reasons for success or failure, and hence develops recommendations for future work.<br />

Terminology used in biological control research can, at times, be ambiguous. The<br />

simple term "parasite" can be particularly troublesome and as an aid to readers its usage<br />

is defined more precisely. More effort has been devoted to entomopathogens during<br />

the current review period and in some cases both microbial parasites and insect<br />

parasites have been examined from the same host insect. To avoid confusion when<br />

referring to these entomogenous and entomophagous parasites, the latter have been<br />

termed parasitoids or insect parasites rather than just parasites. A parasitoid,<br />

following the definition <strong>of</strong> DeBach, is an insect that develops as a larva on or in a single<br />

host insect from eggs laid on, in or near the host; it usually consumes most <strong>of</strong> the host's<br />

tissues and it kills the host. Phytophagous insects that feed on plants regarded as<br />

weeds are thus termed parasites, whereas those feeding on crop plants, including trees,<br />

are referred to as pests (emphasizing their relationship with man rather than with the host<br />

plant).<br />

The term ''biological control" also requires some explanation. In the present context<br />

the term refers to an applied control strategy that involves the manipulation <strong>of</strong> living<br />

natural enemies for the purpose <strong>of</strong> regulating the abundance <strong>of</strong> pest populations. The<br />

regulation may be short or long term and the manipulation may include release<br />

strategies ranging from spot-introductions to widespread inundation. Introductions <strong>of</strong><br />

sterile individuals are thus included here; but the use <strong>of</strong> semiochemicals, notably<br />

kairomones and pheromones, is included only where these chemicals influence<br />

predator or parasitoid behaviour. Pheromone confusion trials on the pest insect are<br />

excluded since they involve direct chemical manipUlation <strong>of</strong> the pest. The use <strong>of</strong><br />

insect growth regulators is similarly excluded.<br />

The book is divided into three sections which cover agricultural insects, weeds, and<br />

forest insects; chapters within each section are generally listed in the alphabetical<br />

order <strong>of</strong> the insect's or weed's generic name; and each <strong>of</strong> the three sections begins with<br />

an overview <strong>of</strong> the chapters that follow. Addresses <strong>of</strong> contributors are listed in the<br />

Appendix. The book does not include an updated synopsis <strong>of</strong> the organization and<br />

accomplishments <strong>of</strong> Canadian biological control research in a separate section<br />

equivalent to Part IV <strong>of</strong> the 1971 volume. Accomplishments are summarized in the<br />

overview chapter beginning each section and the organizational observations in the<br />

1971 synopsis remain equally valid today without need for further elaboration.


x Gcncrallntroduction<br />

Several changes in the organization <strong>of</strong> Canadian biological control research have<br />

taken place since publication <strong>of</strong> the 1971 book. The Research Institute <strong>of</strong> Agriculture<br />

<strong>Canada</strong> at Belle .. ille, Ontario, which had played a pivotal role in importation and<br />

applied research <strong>of</strong> biological control organisms was closed in 1972 and its staff was<br />

relocated. The Biocontrol Unit was formed in Ottawa by Agriculture <strong>Canada</strong> to coordinate<br />

importations. Material is imported through the Biocontrol Unit, and through<br />

Macdonald College in Quebec, the Forest Pest Management Institute and the<br />

University <strong>of</strong> Guelph in Ontario, and the Regina Research Station <strong>of</strong> Agriculture<br />

<strong>Canada</strong> in Saskatchewan. The Department <strong>of</strong> Fisheries and Forestry became part <strong>of</strong> a<br />

new Department, Environment <strong>Canada</strong>, in 1971. Forestry research is conducted by the<br />

Canadian Forestry Service within this Department; biological control research is still<br />

carried out at the same laboratories as those listed in the last review but establishment<br />

names have been modified to reflect the new organization.<br />

Many people assisted in the preparation <strong>of</strong> this review. Special acknowledgement is<br />

extended to Ms. R. Boyle and Ms. P. Loshak for helping in correcting the text. Dr. F.D.<br />

Bennett, Director, Commonwealth Institute <strong>of</strong> Biological Control, initiated the<br />

preparation, Dr. 0.1. Greathead, Assistant Director and Mr. 0.1. Girling, Information<br />

Officer, assisted in its publication, and Mrs. A.H. Greathead read the pro<strong>of</strong>s and<br />

completed the index.


PART I<br />

BIOLOGICAL CONTROL OF<br />

AGRICULTURAL INSECTS IN<br />

CANADA 1969-80


Chapter 1<br />

Current Approaches to Biological<br />

Control <strong>of</strong> Agricultural Insect Pests<br />

1.S. KELLEHER<br />

This part covers the biological control <strong>of</strong> insect and mite pests <strong>of</strong> crops, fruit trees, and<br />

ornamentals. As in the previous review (Kelleher 1971) it includes inoculation<br />

(introduction), augmentation, and inundation, using beneficial insects, entomopathogens,<br />

and autocidal methods.<br />

Some organizational changes have occurred since the last review. In 1972, the<br />

Research Institute at Belleville, Ontario, was closed and the staff transferred to<br />

Research Stations in other parts <strong>of</strong> <strong>Canada</strong>. The service function, <strong>of</strong> obtaining<br />

infonnation and collections <strong>of</strong> biological control organisms from abroad was transferred<br />

to Ottawa. The programme on the biological control <strong>of</strong> weeds was moved to the Regina<br />

Research Station. An Integrated Pest Control Section was fonned at Winnipeg with<br />

many <strong>of</strong> the scientists from Belleville. Other individuals were stationed at Summerland,<br />

Saskatoon, Harrow, and Ottawa.<br />

The adoption <strong>of</strong> integrated pest management (IPM) systems has placed biological<br />

control in a much more important role than before. At one time it was considered as an<br />

alternative to chemicals i.e. complete control was necessary. Now, however, our<br />

knowledge <strong>of</strong> insect population dynamics has increased, and it is now considered that<br />

even a small additional percentage mortality can have a pr<strong>of</strong>ound effect on the overall<br />

reduction <strong>of</strong> an insect population.<br />

The initial stages <strong>of</strong> an IPM programme involve a reduction in pesticide usage, by<br />

monitoring pest infestations so that chemicals are applied only when necessary. As a<br />

consequence, natural control factors are not so suppressed and are able to exert<br />

additional mortality on the pest population. But there will be a point where very little<br />

more can be obtained from the systems available and additional factors must be added<br />

to attain a greater reduction in pesticide use. This can come in many forms but the<br />

introduction <strong>of</strong> new natural enemies should be an important consideration in many<br />

instances.<br />

Most <strong>of</strong> the studies reported in this part have utilized introduced species <strong>of</strong><br />

parasitoids, or entomopathogens (Bacillus thuringiensis Berliner (B.t.), fungi, viruses).<br />

Other aspects <strong>of</strong> biological control such as mass releases <strong>of</strong> parasitoids e.g. Trichogramma<br />

species, have been largely neglected. However reports from the USSR,<br />

China, and South American countries have attributed considerable successes with this<br />

approach and are prompting some consideration <strong>of</strong> initiating similar work in <strong>Canada</strong>.<br />

Much may depend on the development <strong>of</strong> techniques to provide a large number <strong>of</strong><br />

parasitoids during a relatively short period for any particular insect. The main<br />

emphasis will be on production and storage but consideration should also be given to a<br />

number <strong>of</strong> target pests whose susceptible stages occur at different times.<br />

Increased awareness <strong>of</strong> the value <strong>of</strong> natural enemies has led to research on the<br />

differential effects <strong>of</strong> pesticides on natural enemies. Although pesticides usually affect<br />

a wide spectrum, some taxa are less affected by them than others. A working group <strong>of</strong><br />

the West Palaearctic Regional Section <strong>of</strong> the International Organization for Biological<br />

Control has taken an organized and coordinated approach to this subject. Members <strong>of</strong><br />

this group have standardized the culturing <strong>of</strong> various taxa <strong>of</strong> beneficial insects and are<br />

adopting standardized methods for testing the effects <strong>of</strong> pesticides on these insects in<br />

the laboratory and in the field. Eventually these data will be available as recommendations<br />

3


4 J. S. Kelleher<br />

Table 1<br />

to be placed on the label <strong>of</strong> pesticide containers and ill the bulletins <strong>of</strong> <strong>of</strong>ficial agencies.<br />

The following 24 chapters in this part have been arranged in alphabetical order using<br />

the scientific name <strong>of</strong> the pest species. An evaluation <strong>of</strong> the programme is given in<br />

Table 1. Only 10 <strong>of</strong> the pests were included both in this review and the last one. It can<br />

be concluded that no other information is available on the other target pests. This<br />

indicates the transient nature <strong>of</strong> many <strong>of</strong> the agricultural projects, in contrast to those<br />

on forest insects and weeds.<br />

Evaluation <strong>of</strong> biological control measures against agricultural insect and mite pests<br />

Pest Method Agent<br />

Acyrthosiphon pisum (Harris) Adventitious Predator<br />

Adelphocoris lineolatus (Goeze) Inoculation Parasitoid<br />

Agromyza frontella (Rondani) Inoculation Parasitoid<br />

Artogeia rapae (L.), Trichoplusia<br />

ni (Hilbner), & Plutella Inundative Virus, B.t.<br />

xylostella (L.) Inoculation Parasitoid<br />

Culex pipiens L. Augmentation Planaria<br />

Culiseta inornata (Williston) Augmentation Fungus<br />

Cydia pomonella (L.) Inundation Virus<br />

Delia antiqua (Meigen) Autocidal<br />

Entomoscelis americana Brown··<br />

Euxoa messoria (Harris) Augmentation Virus<br />

Forficula auricularia (L.) Inoculation Parasitoid<br />

Hypera postica (Gyll.) Adventitious Fungus<br />

Lygus spp. Inoculation Parasitoid<br />

Mamestra configurata Walker··<br />

Manduca quinquemaculata (Haworth) Inundation B.t.<br />

Melanopus spp., Camnula pellucida<br />

(Scudder) & other grasshoppers Inundation Protozoa<br />

Musca domestica L. Integration Parasitoid<br />

Oulema melanopus (L.) Inoculation Parasitoid<br />

Phyllonorycter blancardella (Fabricius) Inoculation Parasitoid<br />

Phyllotreta spp. Inoculation Parasitoid<br />

Tetranychus unicae Koch· Integration Predator<br />

Thymelicus lineola (Ochs.) Inundation B.t.<br />

Inundation Virus<br />

Inoculation Parasitoid<br />

Tipula paludosa (Meigen) Inoculation Parasitoid<br />

Trialeurodes vaporariorum (Westwood)· Integration Parasitoid<br />

Measures started during or extending into period 1969 to 1980<br />

·Greenhouse<br />

··Exploratory studies - no. released<br />

•• • Degree <strong>of</strong> Success:<br />

- No control or not yet known if established<br />

+ Slight pest reduction or too early for evaluation <strong>of</strong> control<br />

+ + Local control; distribution restricted or not fully investigated<br />

+ + + Control widespread but local damage occurs<br />

+ + + + Control complete<br />

Degree <strong>of</strong><br />

Success···<br />

+<br />

+<br />

++++<br />

++<br />

+++<br />

+<br />

+<br />

+<br />

+<br />

+++<br />

++++<br />

+++<br />

++++<br />

++++<br />

+<br />

+++<br />

++++<br />

+++<br />

+<br />

++++


Literature Cited<br />

Current approaches to biological control <strong>of</strong> agricultural insect pests 5<br />

In the last Review, the use <strong>of</strong> a nematode, DD-136, was reported (Welch 1971)<br />

against several crop pests. This work was not continued although there are still several<br />

avenues to explore. Hopefully this will be done in the future. It will also be noted from<br />

Table 1 that predators were used in only 2 instances - one adventitiously, the other in<br />

a greenhouse situation. Possibly this group is less favored because the members tend<br />

to be less specific than parasitoids. However for reasons considered above, even<br />

relatively small degrees <strong>of</strong> mortality can effect population reductions on a much larger<br />

scale. But it would also be important to determine if there would be a net increase in pest<br />

mortality or if it would be merely compensatory.<br />

Of the 26 agents used, 11 were by inoculation with parasitoids and 7 by inundation<br />

with pathogens. Inoculations gave high success in 2 cases; others could not be<br />

evaluated as it was too early or recoveries had not yet been made. In contrast, 5 <strong>of</strong> the<br />

7 pests were controlled well by inundation with pathogens. However this method is<br />

primarily for short-term stage control whereas inoculation <strong>of</strong> new natural enemies is a<br />

long-term population control measure.<br />

In many cases, establishment <strong>of</strong> exotic natural enemies may be hampered by the low<br />

numbers available for release e.g. against Adelphocoris lineolatus (Goeze), Phyllotreta<br />

spp., and Thymelicus lineola (Ochs.). Propagation has been difficult or unattainable.<br />

Sometimes mortality has been high in rearing andlor storage. This has required intense<br />

efforts by the CIBC to increase the number <strong>of</strong> individuals obtained. Some new<br />

techniques have been considered e.g. leaving an area undisturbed on the premise that<br />

parasitoid populations would increase. Increasing host populations by the addition <strong>of</strong><br />

their eggs from other areas has also been tried successfully in some cases although the<br />

danger <strong>of</strong> introducing more virulent strains has restricted the use <strong>of</strong> this technique.<br />

In addition to the work reported in the following chapters there have been some<br />

minor efforts which did not warrant separate sections. These are as follows: Mantis<br />

religiosa L. ootheca were collected near Belleville, Ontario, in 1971 and 1972 and<br />

released near St. John's, Newfoundland against various grasshopper species (Acrididae);<br />

no recoveries have been recorded (R. Morris, 1981, personal communication).<br />

Coccinella septempunctata L. was found in northern New Jersey, in 1973 apparently<br />

as an adventitious introduction (Angalet & Jacques 1975). They were redistributed to<br />

various states in subsequent years and to Nova Scotia in 1981. The species had been<br />

reported from various parts <strong>of</strong> Quebec (Larochelle 1979) and was found in New<br />

Brunswick in 1981 (P.W. Schaefer, 1981, personal communication).<br />

As in previous reviews, this publication has been possible only through the cooperation<br />

<strong>of</strong> those who performed the work. In this part. 28 authors have contributed. Their efforts<br />

are gratefully acknowledged. as well as those <strong>of</strong> editors and typists who are too numerous<br />

to name individually.<br />

Angalct. G. W.; 1 acquest R.L. (1975) The establishmcnt <strong>of</strong> Coccinella septempunctata L. in thc continental Unitcd Statcs. USDA Cooperative<br />

Economic Insect Report 25, 883-884.<br />

Kcllchcr. 1.S. (1971) Current approachcs to biological control <strong>of</strong> agricultural insect pests. In: Biological control programmes against insects<br />

and weeds in <strong>Canada</strong> 1959-1968. Commonwealth Institute <strong>of</strong> Biological Control Technical Communication<br />

4.1-3.<br />

Larochellc, A. (1979) Les coleoptercs Coccinellidac du Quebec. Cordulia. Supplement 10. I-Ill.<br />

Wclch. H.E. (1971) Various target species allcmplS with 00-136. In: Biological control programmes against insects and weeds in <strong>Canada</strong><br />

1959-1968. Commonwealth Institute <strong>of</strong> BiologiClll Control TechniCIJI Communication 4. 62-66.


Blank Page<br />

6


Literature Cited<br />

Chapter 2<br />

Acyrthosiphon pisum (Harris), Pea Aphid<br />

(Homoptera: Aphididae)<br />

B.D. FRAZER<br />

Coccinella undecimpunctata L. was first encountered in pea aphid, Acyrthosiphon<br />

piswn (Harris), infested alfalfa in May 1972. Since then it is routinely found on low-growing,<br />

aphid infested plants including cereals, strawberries, and vegetables in numbers<br />

about equal to the predominant native species, C. californica and C. trifasciata. C.<br />

undecimpunctata is one <strong>of</strong> the first coccineJlids to be found in spring. They appear to<br />

be adapted to cooler cJimates than the lower mainland <strong>of</strong> British Columbia as there are<br />

few aphids available when they are first seen. The effect <strong>of</strong> their predation in reducing<br />

spring aphid numbers can greatly reduce the rate <strong>of</strong> increase <strong>of</strong> aphids (Frazer &<br />

Gilbert 1976) and thereby extend the time before pesticides are required. Later in the<br />

season the effects <strong>of</strong> C. undecimpunctata alone are smalJ because losses from the large<br />

number <strong>of</strong> predators in alfalfa, at least, are compensatory (Charnov et al. 1976). If C.<br />

undecimpunctata were added to a field without them, their aphid consumption would<br />

simply result in less for the other competing predators with no net increase in aphid<br />

mortality.<br />

Frazer, B.D.; Gilben. N. (1976) Coccinellids and aphids: a quantitative study <strong>of</strong> the impact <strong>of</strong> adult ladybirds (Coleoptera: CoccineUidae) preying<br />

on field populations <strong>of</strong> pea aphids (Homoptera: Aphididae). Journal <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong><br />

British Columbia 73, 33-56.<br />

Charnov, E.; Frazer, B.D.; Gilbert. N.; Raworth. D. (1976) Fishing for aphid - the exploitation <strong>of</strong> a natural population. JourfllJl <strong>of</strong> Applied<br />

Ecology 13. 379-389.<br />

7


Blank Page<br />

8


\0 C. H. Craig and C. C. Loan<br />

Recommendations<br />

Literature Cited<br />

development at intended liberation sites in western <strong>Canada</strong>. However, 47 P. ade/phocoridis<br />

were released near Clavet, Saskatchewan, latitude 52°N, in alfafa containing a<br />

second and third instar A. Iineo/alllS nymph population <strong>of</strong> 2 per sweep. A total <strong>of</strong> 29 P.<br />

ade/phocoridis and 49 P. rubricol/is were released in a similar situation near Shellbrook.<br />

Saskatchewan. latitude 53°N. Liberations in 1981 included 16 P. ade/phocoridis at the<br />

same site at Saskatoon and 12 <strong>of</strong> the species at Shellbrook. The first attempt to recover<br />

these introduced parasitoids at the liberation sites will be made in 1981.<br />

The introduction and release <strong>of</strong> P. ade/phocoridis and P. rubricol/is against A.<br />

lineo/atus in western <strong>Canada</strong> will continue through 1982 or 1983 when the impact <strong>of</strong> the<br />

programme will be assessed. Because A. Iineo/alllS is an excellent candidate host for<br />

introduced parasitoids. it is recommended that a search be made for other specific<br />

parasitoids <strong>of</strong> A. lineo/atus in Europe to increase the probability <strong>of</strong> successful<br />

establishment in western <strong>Canada</strong>.<br />

Bilewicz·Pawinska. T. (1975) Distribution <strong>of</strong> the insect parasites Peristenus Foerster and Mesochorus Gravenhorst in Poland. Bulletin de<br />

I'Academie Polonaise des Sciences Closse II Serie des Sciences Biologiques 23. 823-827.<br />

Craig. C.H. (1971) Distribution <strong>of</strong> Adelphocoris lineolatus (Heteroptera: Miridae) in western <strong>Canada</strong>. CanadUJn Entomologist 103. 280-81.<br />

Craig, C.H. (1963) The alfatra plant bug, Adelphocoris lineolatus (Goeze) in northern Saskatchewan. Canadian Entomologist 95.6-13.<br />

Loan. C.C. (1965) Life cycle and development <strong>of</strong> Peris/enus pal/ipes (Curtis) (Hymenoptera: Braconidae. Euphorinae) in five mirid hosts in the<br />

Belleville district. Proceedings <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> Ontario (1964) 95. 115-121.<br />

Loan. C.c. (1979) Three new species <strong>of</strong> Peristenus Foerster from <strong>Canada</strong> and western Europe (Hymenoptera: Braconidae. Euphorinae).<br />

Natura/iste Canadian 106.387-391.


Pest Status<br />

Background<br />

Chapter 4<br />

Agromyza frontella (Rondani), Alfalfa<br />

Blotch Leafminer (Diptera:<br />

Agromyzidae)<br />

J.C. GUPPY, D.G. HARCOURT, M.O'c. GUIBORD and<br />

L.S. THOMPSON<br />

The alfalfa blotch leafminer, Agromyza frontella (Rondani), was introduced from<br />

Europe to the north eastern United States during the late 1960s (Miller & Jensen 1970)<br />

and first appeared in <strong>Canada</strong> at St-Armand, Quebec, in 1972 near the United States<br />

border (Harcourt 1973). The insect spread rapidly throughout eastern <strong>Canada</strong> from<br />

Prince Edward Island to Ontario (Thompson 1974, Guppy & Harcourt 1977), and by<br />

1979 it had threatened to become the most important pest <strong>of</strong> alfalfa in the 10 eastern<br />

counties <strong>of</strong> Ontario (Harcourt & Binns 1980) and in the other eastern provinces. It has<br />

continued to spread westward in Ontario and damaging levels were recorded in<br />

Hastings County in 1980 (Guppy 1981).<br />

Hosts include: Medicago, Melilollls, and Trifolium, but alfalfa, Medicago sativa L.,<br />

is apparently preferred (Steyskal 1972, Spencer 1973). The insect overwinters as a<br />

partially developed pupa and the adult emerges about mid-May to oviposit in the firstgrowth<br />

alfalfa. In eastern <strong>Canada</strong> there are three complete generations a year (Guppy<br />

1981), but a few individuals may complete a fourth. The crop is attacked by both the<br />

adults and larvae, but most <strong>of</strong> the damage is done by the larvae which feed on the<br />

mesophyll tissues forming blotch mines each <strong>of</strong> which represents about 27% <strong>of</strong> the<br />

leaflet area.<br />

The crop is attacked throughout the growing season but damage is more severe in<br />

the first and second cuttings; during the past 4 years, the percentage <strong>of</strong> leaflets mined<br />

in eastern Ontario averaged 70% and 58%, respectively.<br />

Seven parasitoid species native.to eastern <strong>Canada</strong> have been reared from larvae <strong>of</strong> A.<br />

frontella, all <strong>of</strong> them are hymenopterous species <strong>of</strong> the family Eulophidae. Diglyphus<br />

websteri (Crawford) and D. begini (Ashmead) were recorded in eastern Ontario and<br />

these two plus D. pulchripes (Crawford) in Quebec. In Prince Edward Island, five<br />

species were found, namely, D. begini, D. intermedius (Girault), Notanisomorpha sp.,<br />

Pnigalio maculipes (Crawford), and Pnigalio n. sp. One larval-pupal parasitoid<br />

Cyrtogaster sp. (Hymentoptera: Pteromalidae) emerged from puparia in Ontario.<br />

Assessment <strong>of</strong> parasitism has shown that the native species are <strong>of</strong> little importance<br />

in suppressing alfalfa blotch leafminer populations. However, a hemipteron species<br />

Nabis ferus (Linnaeus), common to eastern <strong>Canada</strong>, appears to be an important<br />

predator <strong>of</strong> the larvae. At Ottawa, incidence <strong>of</strong> attack increases with each generation,<br />

averaging about 60% in the third.<br />

The Beneficial Insects Research Laboratory, Newark, Delaware, United States<br />

Department <strong>of</strong> Agriculture imported 14 parasitoid species from Europe from 1974 to<br />

1978 for release in the United States (Hendrickson & Barth 1979). By 1978 three<br />

hymenopterous species, Dacnusa dryas (Nixon) (Braconidae), Chrysocharis punctifacies<br />

Delucchi (Eulophidae), and Miscogaster hortensis Walker (Pteromalidae), were<br />

11


12 J. C. Guppy, D. G. Harcourt, M. O'c. Guibord and L. S. Thompson<br />

established at the Delaware field plots. These were recolonized in other northeastern<br />

states, where they became established (Hendrickson & Barth 1979, and personal<br />

communication), and imported for release into eastern <strong>Canada</strong>. A fourth species,<br />

Diglyphus isaea (Walker) (Eulophidae), was shipped to Quebec for release in 1975;<br />

establishment was not successful in either country apparently because it crossed with<br />

native species <strong>of</strong> Diglyphus and further attempts <strong>of</strong> establishment were abandoned<br />

(R.M. Hendrickson 1976, personal communication).<br />

Releases and Recoveries D. dryas, C. punclifacies, and M. hortensis are solitary endoparasitoids. The host is<br />

attacked during the larval stage and the parasitoids complete their development during<br />

Table 2<br />

the pupal stage.<br />

All releases in <strong>Canada</strong> were <strong>of</strong> adults, laboratory reared in Newark, Delaware, or<br />

field collected there or in other northeastern States by cooperators and shipped to<br />

<strong>Canada</strong> directly to release points or for distribution through Agriculture <strong>Canada</strong>,<br />

Research Branch, Biological Control Unit, Research Program Service. Release and<br />

recovery data are given in Table 2. D. dryas, the only species thus far recovered,<br />

Releases and recoveries <strong>of</strong> parasitoids against the alfalfa blotch leafminer, Agromyza<br />

fronlella (Rondani), in <strong>Canada</strong>.<br />

Releases<br />

Species and Province Year Number Y car <strong>of</strong> recovery<br />

Chrysocharis plillctifacies<br />

Prince Edward Island 1978 4<br />

Prince Edward Island 1980 142<br />

Ontario 1980 150<br />

Quebec 1979 249<br />

Quebec 1980 90<br />

DaClIlIsa dryas<br />

Prince Edward Island 1977 68<br />

Prince Edward Island 1978 70<br />

Prince Edward Island 1979 546 1980<br />

Ontario 1979 586 1979<br />

Quebec 1979 16<br />

Quebec 1980 378 1981<br />

Dig/ypllUs isaea<br />

Quebec 1975 900<br />

Miscogasler IlOrlellsis<br />

Prince Edward Island 1978 32<br />

appears to be firmly established in eastern <strong>Canada</strong>. In eastern Ontario, it was reared<br />

from third generation pupae in the year <strong>of</strong> release, 1979, and from each generation in<br />

1980 after overwintering successfully in diapausing pupae. It was also collected in<br />

sweepnet samples throughout the growing season in 1980 at Ottawa.


Blank Page<br />

14


Pest Status<br />

Use <strong>of</strong> Bacillus thuringiensis<br />

Chapter 5<br />

Field Plot Introductions <strong>of</strong> Viruses<br />

Artogeia rapae (L.), Imported<br />

Cabbageworm (Lepidoptera: Pieridae),<br />

Trichoplusia ni (Hiibner), Cabbage<br />

Looper (Lepidoptera: Noctuidae) and<br />

Plutella xylostella (L.), Diamondback<br />

Moth (Lepidoptera: Plutellidae)<br />

R.P. JAQUES and J.E. LAING<br />

Artogeia rapae (L.), the imported cabbageworm, causes damage to cole crops<br />

practically wherever these crops are grown in <strong>Canada</strong>. The pest is quite susceptible to<br />

certain chemical insecticides and to insecticides containing Bacillus thuringiensis<br />

Berliner and, therefore, populations <strong>of</strong> larvae are usually retained at acceptable<br />

densities by recommended control programmes. Trichoplusia ni (Hiibner), the cabbage<br />

looper, is not as widespread but, because it is less susceptible than A. rapae to<br />

registered insecticides, it is considered to be the more important pest <strong>of</strong> cole crops,<br />

especially in southern Ontario. Plutella xylostella (L.) is not considered to overwinter<br />

in <strong>Canada</strong> and thus does not become a pest on cole crops until late in the season.<br />

Formulations <strong>of</strong> B. thuring;ensis (B.t.) containing B. thuringiensis var. kurstaki, a<br />

high-potency strain <strong>of</strong> B.t., were registered in <strong>Canada</strong> in 1972 for use against A. rapae,<br />

T. ni, and P. xylostella on cole crops and other crops. The new formulations (Thuricide®<br />

HPC, Sandoz Crop Protection Inc.; and Dipel® HD-l, Abbott Laboratories Ltd.)<br />

replacing formulations registered previously, protected cole crops equally as well as or<br />

better than did chemical insecticides. These formulations <strong>of</strong> B.I. are now used quite<br />

extensively.<br />

The native parasitoid complex <strong>of</strong> the imported cabbageworm (Michalowicz 1980)<br />

and the diamondback moth (Butts 1979, Bolter 1982) have been studied intensively.<br />

Two species, Apanteles rubecula Marsh. which parasitizes A. rapae and Apanleles<br />

plulellae (Kurdj.) a parasitoid <strong>of</strong> diamondback moth, have been imported and released<br />

in southwestern Ontario.<br />

The granulosis virus <strong>of</strong> A. rapae (ARGV), a naturally occurring virus (Jaques &<br />

Harcourt 1971) that <strong>of</strong>ten causes high mortality in populations <strong>of</strong> larvae <strong>of</strong> A. rapae<br />

late in the growing season, was evaluated by bioassays and plot tests as an introduced<br />

pathogen. In addition, three viruses that kill larvae <strong>of</strong> T. ni were assessed in laboratory<br />

and field studies. Two <strong>of</strong> these viruses, the single-embedded nuclear polyhedrosis<br />

15


16 R. P. Jaques and J. E. Laing<br />

virus <strong>of</strong> T. n; (TNNPV-SEV), the predominant naturally occurring virus <strong>of</strong> T. ni<br />

(Jaques 1970a) , and the nuclear polyhedrosis virus <strong>of</strong> Autographa cali/orn;ca<br />

(ACNPV), a virus that kills T. n; and that closely resembles the multiple-embedded<br />

nuclear polyhedrosis virus <strong>of</strong> T. ni (TNNPV-MEV), are considered as candidates for<br />

development as microbial insecticides for use against T. n;.<br />

Efficacy <strong>of</strong> the viruses <strong>of</strong> A. rapae and T. n; was assessed in 1970-80 using field<br />

plots <strong>of</strong> cabbage (each 24-32 m2) Jaques 1970b, 1972, 1973, 1977; Jaques & Laing<br />

1978).<br />

In these tests:<br />

(1) ARGV (7.5 x 10 12 granular inclusion bodies (GIBlha) applied to foliage 4 to 6<br />

times during the growing season reduced populations by 95 to 100% and protected the<br />

crop as well as did chemical insecticides applied at the recommended rates.<br />

(2) TNNPV-SEV, TNNPV-MEV and ACNPV were about equally effective against<br />

the cabbage looper. Four to 6 applications <strong>of</strong> 1.5 x lOll polyhedral inclusion bodies<br />

(PIB)lha during the growing season reduced populations <strong>of</strong> T. n; on cabbage by about<br />

75% and protected the crop as well as did methomyl or B.t. but not as well as<br />

permethrin applied at recommended rates.<br />

(3) Reduced-dosage mixtures <strong>of</strong> ACNPV, ARGV, andlor B.t. with chlordimeform<br />

were as effective against A. rapae and T. ni as were the microbial or chemical<br />

insecticides applied alone at the full rate (Jaques & Laing 1978). Mixtures <strong>of</strong> the<br />

viruses or B.t. and permethrin at 114 or 112 full rate were equally as effective as<br />

permethrin applied alone in protecting the crop.<br />

(4) The efficacy <strong>of</strong> ACNPV propagated ;n vitro was similar to that <strong>of</strong> ACNPV<br />

propagated ;n vivo.<br />

(5) Because the viruses persist in soil (Jaques 1974a, 1974b), application <strong>of</strong> a high<br />

dosage <strong>of</strong> ARGV and ACNPV when plants were transplanted, suppressed populations<br />

<strong>of</strong> A. rapae and T. n; on cabbage plants and lessened the frequency <strong>of</strong> use <strong>of</strong> insecticides<br />

later in the season.<br />

Predation and Parasimm The native parasitoid complex <strong>of</strong> the imported cabbageworm and the diamondback<br />

moth have been studied intensively (Butts 1979, Michalowicz 1980, Bolter 1982).<br />

Major mortality factors for A. rapae at Cambridge, Ontario during 1977-78 were<br />

Apanteles glomeratus (L.), Phryxe vulgaris (Fallen), Compsilura concinnata (Meigen),<br />

and Pteromalus puparum L. Major changes in the population occurred during the<br />

pupal stage and the key factor responsible for these changes was the pupal parasitoid, P.<br />

puparum. A. glomeratus which parasitizes the larvae <strong>of</strong> A. rapae also contributed<br />

substantially to the total mortality. However, rates <strong>of</strong> parasitism by both species <strong>of</strong><br />

parasitoids only reached high levels late in the growing season and did not prevent A.<br />

rapae from reaching densities which were injurious to the cabbage crop (Michalowicz<br />

1980).<br />

Major mortality factors <strong>of</strong> the diamondback moth larvae in Ontario are the larval<br />

parasitoids, M;croplitis plutellae (Mues.) and Diadegma insulare (Cress.), and the<br />

pupal parasitoid, D;adromus subti/icornis (Grav.) (Butts 1979, Bolter 1982). In three<br />

<strong>of</strong> four years in which parasitoids were monitored from 1977-81 in Cambridge,<br />

Ontario, D. insulare was the dominant parasitoid while D. subtilicornis was the<br />

dominant parasitoid in one <strong>of</strong> the four years.<br />

In a study <strong>of</strong> the competition between the larval parasitoids D. insulare and M.<br />

plutellae (Bolter 1982), it was shown that female D. insulare produced an average <strong>of</strong><br />

814 eggs in a lifespan <strong>of</strong> 26 days at 23°C, while female M. plutellae produced an average<br />

<strong>of</strong> 318 eggs in a lifespan <strong>of</strong> 20 days at 23°C. D. insulare avoided superparasitism and<br />

multiple parasitism <strong>of</strong> hosts already parasitized by M. plutellae. M. plutellae generally


Artogeia rapae (L.). 17<br />

avoids superparasitism. but cannot detect eggs <strong>of</strong> D. insulare in the host for at least 12<br />

hours after they were oviposited. However, the first-instar larva <strong>of</strong> M. plutellae was<br />

shown to be intrinsically superior to the first-instar larva <strong>of</strong> D. insulare.<br />

The native parasitoid complex <strong>of</strong> the diamondback moth does not provide adequate<br />

control <strong>of</strong> diamondback moth on Brussels sprouts late in the season when conditions<br />

favour an increase in the host population.<br />

Parasitoid Introductions Apanteles rubecula a larval parasitoid <strong>of</strong> A. rapae was introduced from British<br />

Columbia to Richmond, Ontario by Harcourt in 1970 and 1971. Establishment <strong>of</strong> the<br />

parasitoid was in doubt. However, four specimens <strong>of</strong> A. rubecula were recovered from<br />

a garden plot in Ottawa in 1982. A small number <strong>of</strong> A. rubecula. originally obtained<br />

from British Columbia, were released in Guelph, Ontario (40 99 and dd in 1978, 85 99<br />

and 120 dd in 1979), Harrow, Ontario (180 99 and 120 dd in 1978) and Cambridge,<br />

Ontario (165 99 and 175 o'd in 1979). Although pupae <strong>of</strong> A. rubecula were observed in<br />

the field in the fall during the year <strong>of</strong> each release, no A. rubecula have been recovered<br />

from the release areas during 1980-82.<br />

A second strain <strong>of</strong> A. rubecula obtained from north eastern China in 1981 was<br />

released at Guelph during 1982, but establishment <strong>of</strong> the parasitoids is unknown at the<br />

time <strong>of</strong> this report.<br />

Apanteles plutellae (Kurdj.) a larval parasitoid <strong>of</strong> the diamondback moth, native to<br />

Europe was imported to Guelph in 1981 from the CIBC laboratory in Trinidad. It is<br />

presently in culture at the University <strong>of</strong> Guelph and will be released in Guelph during<br />

1982.<br />

Host-Parasitoid-Pathogen Interactions<br />

The host-parasitoid-pathogen interactions between the parasitoid Apanteles glomeratus,<br />

a granulosis virus (GV), and their lepidopteran host Artogeia rapae larvae<br />

depended on the length <strong>of</strong> time between oviposition by the parasitoid and time <strong>of</strong><br />

ingestion <strong>of</strong> the virus by the host. Parasitoids did not survive in hosts which ingested<br />

virus on or before the fifth day post-parasitism. Percentage survival <strong>of</strong> parasitoids<br />

gradually increased from 26% when oviposition by parasitoids occurred six days prior<br />

to GV infection, to 90% when oviposition by parasitoids occurred nine days prior to GV<br />

infection. Survival <strong>of</strong> parasitoids in non-virus treated hosts remained consistently<br />

above 95%. The degree <strong>of</strong> host infection was strongly correlated with survival <strong>of</strong><br />

parasitoids: the heavier the host's GV infection, the fewer parasitoids that emerged<br />

from the host to pupate and that emerged from the pupae as adults (Levin et al. 1981).<br />

After prior oviposition in GV-infected A. rapae larvae, A. glomeralUS transmitted<br />

GV to healthy A. rapae larvae 73% <strong>of</strong> the time. Ninety per cent <strong>of</strong> parasitoids<br />

transmitted GV to at least one recipient larva. A. glomeralus which developed in and<br />

emerged from GV-infected A. rapae larvae, transmitted GV to 84% <strong>of</strong> the larvae in<br />

which they initially oviposited. Discrimination between healthy and diseased A. rapae<br />

larvae by female A. glomeralus was not observed. This lack <strong>of</strong> discrimination is likely<br />

to contribute to the dissemination <strong>of</strong> GV from host to host by A. glomeralUS in the field<br />

and to mortality <strong>of</strong> the parasitoid's progeny. Thus, the parasitoid is at a disadvantage<br />

when competing with GV in the field for hosts (Levin el al. 1979, 1983).


Pest Status<br />

Background<br />

Chapter 6<br />

Culex pipiens L., Northern House<br />

Mosquito (Diptera: Culicidae)<br />

1.A. GEORGE<br />

The northern house mosquito, or rain barrel mosquito, Culex pipiens L., has long been<br />

known to live in close proximity to humans in southern Ontario (Wood el al. 1979).<br />

Larvae occur most commonly in stagnant water polluted to some degree with varying<br />

amounts <strong>of</strong> decaying organic matter such as vegetation or sewage. C. pipiens larvae<br />

are usually associated with, or follow the earlier occurring larvae <strong>of</strong> Culex resluans<br />

Theobald. Adult populations <strong>of</strong> C. pipiens reach a peak in August, and though<br />

breeding continues until October, females begin entering diapause in late August and<br />

will not take a blood meal until the following spring (Madder el al. 1980, Madder 1981).<br />

Mated females spend the winter in any protected shelter available to them. About mid­<br />

May females seek their first blood meal and deposit the first egg rafts during the third<br />

week <strong>of</strong> May at Guelph (Madder el al. 1980). Thoughfemales prefer the blood <strong>of</strong> birds<br />

to that <strong>of</strong> mammals (Hayes 1961), they will, when abundant, invade homes and feed on<br />

humans.<br />

C. pipiens has been implicated in the spread <strong>of</strong> St. Louis encephalitis (SLE) and it,<br />

along with Culex resluans (Madder el al. 1980), is assumed to be the vector <strong>of</strong> this<br />

virus in eastern North America (Wood el al. 1979). The first recorded epidemic <strong>of</strong> SLE<br />

occurred in southern Ontario in 1974-75 and stimulated investigations on surveillance<br />

for,and abatement <strong>of</strong> this insect (Mahdy el al. 1979, Helson el al. 1980).<br />

Four years (1977 -80) <strong>of</strong> identifying mosquito larvae by the author for the Middlesex­<br />

London District Health Unit, London, Ontario survey crews, has helped to establish<br />

catch basins, designed to drain away surface water, as major breeding sites for C.<br />

pipiens in this area. In 1979 catch basins accounted for 96% <strong>of</strong> all breeding sites found<br />

by the survey crew (Pringle el al. 1978). Moreover, on randomly selected streets in<br />

Brampton, Clarkson, Streetsville, and Port Credit, 70% <strong>of</strong> the catch basins examined<br />

contained Culex larvae (Hill, 1978, personal communication).<br />

Catch basins not only provide water high in organic matter throughout the summer,<br />

but, in addition, the underground pipes into which they drain provide ideal protected<br />

sites for the females to spend the winter months. In addition, the street locations <strong>of</strong><br />

most catch basins puts them in close proximity to human dwellings. Finally, some<br />

catch basins form a part <strong>of</strong> many farm drainage systems. These also support Culex<br />

larvae. Thus, the drainage systems, established throughout southern Ontario to<br />

remove surface water, provide ideal breeding conditions for C. pipiens and C.<br />

resluans, the suspected vectors <strong>of</strong> SLE.<br />

The planarian flatworm, Dugesia ligrina (Girard) (Tricladida, Turbellaria), has been<br />

known to be a predator <strong>of</strong> mosquitoes since Lischetti (1919) reported it to consume<br />

mosquito larvae (Jenkins 1964). A larger species, Dugesia dorolocephala (Woodworth),<br />

has recently been shown also to act as an effective predator <strong>of</strong> mosquito eggs, larvae,<br />

and pupae in California (Legner & Medved 1972, 1974, Medved & Legner 1974, Yu &<br />

Legner 1976). Small Merosloma species, under 1 mm in length, <strong>of</strong> planaria in the order<br />

Rhabdocoelida, kill mosquito larvae in the rice fields <strong>of</strong> California (Case & Washino<br />

1979).<br />

19


Recommendations<br />

Acknowledgement<br />

Literature Cited<br />

Culex pipiens L.. 21<br />

basins can readily be inoculated with reared planaria or more simply by transferring<br />

floats with planaria to new catch basins.<br />

Research on the abatement <strong>of</strong> Culex spp. in catch basins is required to enable intelligent<br />

control measures to be implemented if Culex-transmitted SLE virus reappears in<br />

Ontario. A study <strong>of</strong> the ability <strong>of</strong> predaceous planaria to survive year round in catch<br />

basins is under way. A search for Merostoma species should be conducted in Ontario.<br />

If planaria fail, the next most promising control in catch basins appears to be a<br />

means <strong>of</strong> slow release <strong>of</strong> B.t.;' or methoprene.<br />

This research was assisted by the Ontario Ministry <strong>of</strong> the Environment through the<br />

Ontario Pesticide Advisory Committee.<br />

Case. T.; Washino. R.K. (1979) Flatwonn control <strong>of</strong> mosquito larvae in rice fields. Science 206. 1412-1414.<br />

George. J.A. (1978) The potential <strong>of</strong> a local planarian. Dugesia tigriM (Tricladida. Turbellaria). for the control <strong>of</strong> mosquitoes in Ontario.<br />

Proceedings <strong>of</strong> Ihe EntomologicDl <strong>Society</strong> <strong>of</strong> Ontario 109.65-69.<br />

Hayes. R.O. (1961) Host preferences <strong>of</strong> Culiseta melanura and allied mosquitoes. Mosquito News 21. 179-187.<br />

Helson. B. V.; Surgeoner. G.A.; Wright. R.E. (1980) The seasonal distribution and species composition <strong>of</strong> mosquitoes (Diptera: Culicidae)<br />

collected during a St. Louis Encephalitis surveillance program from 1976 to 1978 in southwestern Ontario.<br />

Canadian Entomologist 112. 865-874.<br />

Jenkins. D.W. (1964) Pathogens. parasites and predators <strong>of</strong> medically important anhropods- Annotated list and bibliography. Bulletin <strong>of</strong><br />

the World Heallh Organization Supplement to vol 30.<br />

Legner. E.F.; Medved. R.A. (1972) Predators investigated for the biological control <strong>of</strong> mosquitoes and midges at the University <strong>of</strong> California.<br />

Riverside. Proceedings <strong>of</strong> the California Mosquito Control Associotion 40. 109-111.<br />

Legner. E.F.; Medvcd. R.A. (1974) laboratory and small-scale field experiments with planaria (Tricladida. Turbellaria) as biological<br />

mosquito control agents. Proceedings <strong>of</strong> the Califomio Mosquito Control Association 42. 79-80.<br />

Uschelli. A.B. (1919) Un venne del genero Planoria. enemigo natural de las larvas del mosquito. Physis. Buenos Aires. 4.591-595.<br />

Madder. D.J. (1981) Biological studies on Culex pipiens L. and Culex restuans Theo. (Diptera. Culicidae) in southern Ontario. Ph.D. Thesis.<br />

University <strong>of</strong> Guelph. 125 pp.<br />

Madder. D.J.; MacDonald. R.S.; Surgeoner. G.A.; He/son. B.V. (1980) The use <strong>of</strong> oviposition activity to monitor populations <strong>of</strong> Culex<br />

pipiens and Culex restuans (Diptera: Culicidae). Canadian Entomologist 112. 1013-1017.<br />

Mahdy. M.S.; Spence. L.; Joshua. J.M. (Eds) (1979) Arboviral encephalitides in Ontario with special reference to St. Louis encephalitis.<br />

Ontario Ministry <strong>of</strong> Health, 364 pp.<br />

Medved. R.A.; Legner. E.F. (1974) Feeding and reproduction <strong>of</strong> the planarian Dugesia dorotocephala (Woodworth). in the presence <strong>of</strong><br />

Culex peris Speiser. Environmental Entomology 3, 637-641.<br />

Pringle. T.; Ryan. S.; Smith. L. (1978) Middlesex County mosquito·bome encephalitis control programme (1978). Report <strong>of</strong> Middlesex­<br />

London District Health Unit, London. Ontario. 36 pp.<br />

Wood. D.M.; Dang. P.T.; Ellis. R.A. (1979) The insects and arachnids <strong>of</strong> <strong>Canada</strong>. Part 6. The mosquitoes <strong>of</strong> <strong>Canada</strong>. Diptera: Culicidae.<br />

Agriculture <strong>Canada</strong> Publication 1686. 390 pp.<br />

Yu. Hyo-sok; Legner. E.F. (1976) Regulation <strong>of</strong> aquatic Diptera by planaria. Entomophaga 21. 3-12.


Blank Page<br />

22


Pest Status<br />

Background<br />

Host Specificity<br />

Chapter 7<br />

Culiseta inornata (Williston), a Mosquito<br />

(Diptera: Culicidae)<br />

J .A. SHEMANCHUK, H.C. WHISLER and S.L. ZEBOLD<br />

Culiseta inomata (Williston) is a very common mosquito species in the Prairie<br />

Provinces <strong>of</strong> <strong>Canada</strong>. It is found not only in the open country but also in the aspen<br />

grove regions. Females <strong>of</strong> this species pass the winter as fertilized females in rodent<br />

burrows, caves, root cellars, rock piles, and abandoned buildings. Generally, the<br />

species does not appear in large numbers until late July. There are several generations<br />

in a season. C. inomata is a common pest <strong>of</strong> livestock and occasionally <strong>of</strong> man, and is<br />

a vector <strong>of</strong> western equine encephalomyelitis and California encephalitis (Morgante &<br />

Shemanchuk 1967, Shemanchuk & Morgante 1967).<br />

Concern about the environmental contamination with pesticides has led many<br />

countries to examine the use <strong>of</strong> biological agents for the control <strong>of</strong> mosquitoes. An<br />

aquatic fungus, Coelomomyces psorophorae Couch (Chytridiomycetes: Blastocladiales),<br />

parasitic on larvae <strong>of</strong> C. inomata was discovered in the irrigated areas <strong>of</strong> southern<br />

Alberta in 1956 and has been found there every year since. Larval mortalities up to<br />

80% have occurred in some <strong>of</strong> the breeding ponds. This fungus was also found<br />

infecting C. inornata larvae in the irrigated areas <strong>of</strong> Saskatchewan, which indicates a<br />

wide geographic distribution.<br />

The persistence <strong>of</strong> this fungus in the irrigated areas <strong>of</strong> southern Alberta and<br />

Saskatchewan and the ability to colonize the host species <strong>of</strong> mosquitoes in the<br />

laboratory prompted us to study further this particular host-parasite combination.<br />

Laboratory experiments resulted in the discovery <strong>of</strong> a previously unknown obligate<br />

alternate host the copepod, Cyclops vernalis, in the life-cycle <strong>of</strong> Coelomomyces<br />

psorophorae (Whisler et al. 1974, 1975). In the life-cycle, the zygote seeks out and<br />

attaches to the intersegmental membranes <strong>of</strong> the mosquito larva and penetrates the<br />

cuticle leading to the development <strong>of</strong> hyphal bodies, mycelium, and ultimately thickwalled<br />

resistant sporangia in the coelom <strong>of</strong> the larva, which kill the larva. Under<br />

appropriate conditions, these sporangia release zoospores <strong>of</strong> opposite mating type<br />

which infect the alternate host, Cyclops vernalis. Each zoospore develops into a<br />

thallus and eventually gametangia. Gametes <strong>of</strong> opposite mating type fuse either in or<br />

outside <strong>of</strong> the copepod to form the mosquito-infecting zygote.<br />

The discovery <strong>of</strong> the life-cycle led to the development <strong>of</strong> a successful laboratory<br />

procedure for the propagation <strong>of</strong> the fungus and to a better understanding <strong>of</strong> its basic<br />

biology.<br />

With the reliable technique which was developed for consistent and significant<br />

infections in C. inornata in the laboratory, host specificity studies <strong>of</strong> Coelomomyces<br />

psorophorae were possible. The possibility <strong>of</strong> infecting Aedes vexans (Meigen)<br />

received particular attention because larvae <strong>of</strong> this mosquito with light infections <strong>of</strong><br />

Coelomomyces psorophorae have occasionally been found in field sites where heavily<br />

infected larvae <strong>of</strong> C. inornata were also found. Larvae <strong>of</strong> A. vexans, A. dorsalis<br />

23


24 J. A. Shemanchuk. H. C. Wisler and S. L. Zebold<br />

Field Trials<br />

Recommendations<br />

Literature Cited<br />

(Meigen), and A. flavescens (Muller) were exposed to zygotes <strong>of</strong> Coelomomyces<br />

psorophorae, but, despite repeated attempts, infections by this strain were never<br />

attained. C. inornata larvae in the same experiment were infected (Zebold et al. 1979).<br />

These results seemed to support the prevailing opinion that Coelomomyces is a<br />

highly specific parasite. In further tests, however, infection was achieved in Culex<br />

pipiens L., C. quinquefasciatus Say, A. aegypti (L.), A. sierrensis (Ludlow), A.<br />

triseriatus (Say), and C. inornata (Louisiana strain).<br />

A. aegypti larvae were more susceptible than the other species, but maximum<br />

resistant sporangia production was consistently best in the original host species, C.<br />

inornata.<br />

In 1979, a preliminary field trial was conducted at 8-Mile Lake near Lethbridge using<br />

laboratory-cultured Coelomomyces psorophorae. Six artificial pools were established.<br />

One pool was inoculated with zoospores and copepods, one received copepods that<br />

were infected in the laboratory and were ready to release the zygotes, two pools<br />

received zygotes from laboratory-infected copepods, and two pools that received no<br />

fungal inoculum served as controls. Into each <strong>of</strong> the pools, first instar laboratoryreared<br />

larvae <strong>of</strong> C. inornata were introduced and exposed for 24 hours after which<br />

they were brought back to the laboratory for rearing to final instar and for observation<br />

for infection.<br />

The results from this test confirmed our field observation that the fungus is an<br />

effective killer <strong>of</strong> C. inomata larvae. The fungus was most effective when introduced<br />

into mosquito habitats in the zygote stage and produced a 64% kill.<br />

These results further indicate that this fungus will have practical application in<br />

controlling C. inornata as a short-term microbial pesticide and as a long-term inoculum<br />

in mosquito-breeding habitats where the fungus does not occur naturally.<br />

(1) Expand research on the design and development <strong>of</strong> techniques for large-scale<br />

production and long-term storage <strong>of</strong> Coelomomyces psorophorae.<br />

(2) Evaluate Coelomomyces psorophorae as a short-term microbial pesticide and as a<br />

long-term inoculum in mosquito-breeding habitats in the irrigated areas and in areas<br />

where the fungus does not occur.<br />

(3) Continue searching for Coelomomyces infections in other species <strong>of</strong> mosquitoes<br />

and define their life-cycles.<br />

(4) Search for other aquatic fungi that might be domesticated and manipulated for the<br />

control <strong>of</strong> mosquitoes.<br />

Morgante. 0.; Shemanchuk. J.A. (1967) Virus <strong>of</strong> California Encephalitis Complex; isolation from Cuiisela inornata. Science 157. 692-693.<br />

Shemanchuk. J.A.; Morgante. O. (1967) Isolation <strong>of</strong> Western Encephalitis virus from mosquitoes in Albena. Canadian JoumJll <strong>of</strong><br />

Microbiology 14.1-5.<br />

Whisler. H.C.; Zebold; S.L.; Shemanchuk. J.A. (1974) Alternate host for mosquito parasitic Coelomomyces. Nature 251. 715-716.<br />

Whisler. H.C.; Zebold. S.L.; Shemanchuk. J.A. (1975) Life history <strong>of</strong> Coelomomyces psorophorae. Proceedings <strong>of</strong> the National Academy<br />

<strong>of</strong> Sciences 72. 693-696.<br />

Zcbold. S.L.; Whisler H.C.; Shemanchuk. J.A.; Travland. L.B. (1979) Host specilicity and penetration in mosquito pathogen Coe/omomyces<br />

psorophorae. Canadian Journal <strong>of</strong> Botany 57. 2766-2nO.


Pest Status<br />

Background<br />

Orchard Tests<br />

Table 3<br />

Chapter 8<br />

Cydia pomonella (L.), Codling Moth<br />

(Lepidoptera: Tortricidae)<br />

R.P. JAQUES and J.E. LAING<br />

The codling moth, Cydia pomonella (L.), is a major pest <strong>of</strong> apple in most areas <strong>of</strong><br />

<strong>Canada</strong> in which the crop is grown. Several chemical insecticides, notably phosmet,<br />

azinphos-methyl, and permethrin, are effective against the pest. Nevertheless, protection<br />

<strong>of</strong> apples against the codling moth is a major concern in development <strong>of</strong> integrated<br />

pest management programmes for apples, not only because <strong>of</strong> the importance <strong>of</strong> the<br />

pest, but also because the broad-spectrum insecticides used to control it are detrimental to<br />

the predaceous and parasitic arthropods in orchards.<br />

A granulosis virus discovered by Tanada (1964) has been shown to be highly infectious<br />

for C. pomonella larvae in laboratory tests (Keller 1973, Laing & Jaques 1981). The<br />

virus was quite effective against the pest in apple orchard tests in Germany (Huber &<br />

Dickler 1977), reducing damage 10 fruit by C. pomonella by as much as 89%. Falcon el<br />

al. (1968) obtained similar crop protection in tests in California but the virus was less<br />

effective in orchard tests in Ohio (Sheppard & Stairs 1976).<br />

The granulosis virus <strong>of</strong> C. pomonella (CPGV) was applied to apple trees at four<br />

locations in <strong>Canada</strong> between 1974 and 1980 in orchard tests on effectiveness <strong>of</strong> the<br />

virus against the codling moth (Jaques el al. 1977, 1981).<br />

Seven-tree plots were sprayed at rales <strong>of</strong> 1 x 10 9 to 3 X 10 10 GIB <strong>of</strong> CPGV or 0.5 g<br />

azinphos-methyl per litre <strong>of</strong> spray in a 4-year test (1975-78) carried out at the<br />

Research Station at KentviJIe, Nova Scotia. A summary <strong>of</strong> results (Table 3) indicated<br />

that CPGV was not as effective as azinphos-methyl in protecting apples against<br />

damage by larvae <strong>of</strong> C. pomonella. Total damage to picked and dropped apples<br />

harvested from plots treated with CPGV was reduced by an average <strong>of</strong> 95.5%.<br />

Injury to apples by larvae <strong>of</strong> Cydia pomonella (L.) in a 4-year orchard test at KentviJIe.<br />

Nova Scotia, on efficacy <strong>of</strong> a granulosis virus·<br />

Material<br />

Virus<br />

Azinphos-methyl<br />

Check<br />

Injury by C. pomonellallOO apples<br />

Picked fruit Picked + Dropped fruit<br />

SE" DE-· SE DE<br />

1.6 1.0 1.7 1.6<br />

0.1 0.1 0.2 0.4<br />

1.2 3.0 1.3 12.1<br />

·Tests were carried out in cooperation with C.R. MacLellan and K.H. Sanford,<br />

Research Station, Agriculture <strong>Canada</strong>. Kentville, N.S.<br />

··SE and DE refer 10 "shallow-entry" and "deep-entry" injury, respectively.<br />

25


26 R. P. Jaques and J. E. Laing<br />

Table 4<br />

Evaluation <strong>of</strong> Control Attempts<br />

The granulosis virus was applied to plots in apple orchards at the Research Station,<br />

Summerland, British Columbia, in 1975 and 1976; at the Research Station, Vineland<br />

Station, Ontario, in 1974, 1977, and 1978; and at the University <strong>of</strong> Guelph, Guelph,<br />

Ontario, in 1976-80; Table 4 lists data from some <strong>of</strong> the tests.<br />

Injury to apples by larvae <strong>of</strong> Cydia pomonella (L.) in orchard tests on efficacy <strong>of</strong> C.<br />

pomonella granulosis virus'.<br />

Materialllitre Applica.tionsl Injury by c.!omonellallOO<br />

<strong>of</strong> spray generatIon apples picke at harvest<br />

1st 2nd SE·· DE**<br />

Summerland, British Columbia<br />

1975<br />

CPGV 3.2-7.3 x lO9 GIB 2 2 55 32<br />

Check 0 0 22 71<br />

Vineland Station, Ontario<br />

1974<br />

CPGV 3.5-6.5 x 10 9<br />

GIB 3 2 21 5<br />

Phosmet 3.1 g 2 1 2 I<br />

Check 0 0 5 22<br />

Guelph, Ontario<br />

1979<br />

CPGV 3.3xlO'o GIB 3 3 9 2<br />

Check<br />

1979<br />

0 0 11 14<br />

CPGV 3.6xlO'o GIB 2 2 11 2<br />

3 4 8 0<br />

Check 0 0 7 11<br />

·Tests were carried out in cooperation with M.D. Proverbs, Research Station<br />

Summerland, British Columbia, and E.A.C. Hagley and R. Trottier, Research<br />

Station, Vineland Station, Ontario.<br />

"SE and DE refer to "shallow-entry" and "deep-entry" injury. respectively.<br />

Total damage to apples in plots treated with CPGV was not reduced appreciably in<br />

many <strong>of</strong> the tests. Deep-entry (D.E.) damage was reduced by 55% to 100%, approaching<br />

the reductions noted in plots treated with a chemical insecticide. It is noteworthy that<br />

shallow entries (stings) (S.E.) were higher in apples from plots treated with the virus<br />

than in nontreated apples in some tests, e.g. 1975 Summerland test, presumably<br />

because larvae infected by the virus died after they had entered the apple.<br />

The 1979 test at Guelph indicated superior crop protection with increased numbers<br />

<strong>of</strong> applications <strong>of</strong> the virus. In a plot test in 1980 at Guelph, CPGV was applied 2 or 4<br />

times for each <strong>of</strong> the two generations <strong>of</strong> C. pomonella larvae; however, crop<br />

protection was not increased by the higher frequency <strong>of</strong> application.<br />

The granulosis virus <strong>of</strong> C. pomonella did not protect the crop as well as did chemical<br />

insecticides in these tests. However, the virus is considered to have substantial<br />

potential in integrated pest management because it is specific for the target insect.


Cydia pomollella (L.). 27<br />

Further studies will emphasize timing and dosages <strong>of</strong> application, persistence <strong>of</strong> the<br />

virus in the orchard habitat, and application <strong>of</strong> the virus in combination with low<br />

dosages <strong>of</strong> chemical insecticides as means <strong>of</strong> increasing effectiveness.<br />

Literature Cited<br />

Falcon. L.A.; Kane. W.R.; Bethell. R.S. (1968) Preliminary evaluation <strong>of</strong> the granulosis virus for control <strong>of</strong> the codling moth. Journal <strong>of</strong><br />

Economic Entomology 61. 1208-1213.<br />

Huber. J.; Dickler. E. (1977) Codling moth granulosis: Its efficacy in the field in comparison with organophosphorous insecticides. Journal <strong>of</strong><br />

Economic EnlOmology 70. 557-561.<br />

Jaques. R. P.: Maclellan. C. R.; Sanford. K. H.; Proverbs. M.D.; Hagley. E.A.C. (19n) Preliminary orchard tests on control <strong>of</strong> codling moth<br />

larvae by a granulosis virus. Canadian Entomologist 109. 1079-1081.<br />

Jaques. R.P.; Laing. J.E.; Maclellan. C.R.; Proverbs. M.D.; Sanford. K.H.; Trottier. R. (1981) Apple orchard tests on efficacy <strong>of</strong> the<br />

granulosis virus <strong>of</strong> the codling moth. Laspeyresia pomonella (Lepidoptera: Olethreutidae). Ento·<br />

mophaga 26. 111-117.<br />

Keller. S. (1973) Mikrobiologische Bekiimpfung des Apfelwicklers Laspeyresia pomonella (L.) (= Carpocapsa pomonel/Q) mit spezifischen<br />

Granulosisvirus. Zeit.schrift fiir angewandte Entomologie 73. 137-181.<br />

Laing. D.R.; Jaques. R.P. (1981) Codling moth: techniques for rearing larvae and bioassaying granulosis virus. Joumal <strong>of</strong> Economic<br />

Entomology 73. 831-853.<br />

Sheppard. R.E.; Stairs. G.R. (1976) Effects <strong>of</strong> dissemination <strong>of</strong> low dosage levels <strong>of</strong> a granulosis virus in populations <strong>of</strong> the codling moth.<br />

Jou",al <strong>of</strong> Economic Entomology 69, 583-586.<br />

Tanada. Y. (1964) A granulosis virus <strong>of</strong> the codling moth. Carpocapsa pomonella (Linnaeus) (Olethreutidae, Lepidoptera). Journal <strong>of</strong> Insect<br />

Pathology 6.378-380.


Blank Page<br />

28


Chapter 9<br />

Delia antiqua (Meigen), Onion Maggot<br />

(Diptera: Anthomyiidae)<br />

F.L. McEWEN<br />

The onion maggot, Delia anliqua (Meigen), is the most destructive pest <strong>of</strong> onions<br />

throughout the temperate regions <strong>of</strong> the northern hemisphere and in <strong>Canada</strong> causes<br />

losses <strong>of</strong> 50-90% (Finlayson 1959). Effective control was not achieved until the late<br />

1940s when the cyclodiene insecticides became available. These, when applied in the<br />

furrow with the seed, provided excellent control <strong>of</strong> the first generation and additional<br />

treatments <strong>of</strong> DDT against the second and third generation adults reduced damage to a<br />

minimum. Resistance to these insecticides developed and the cyclodienes were<br />

replaced about 1960 by such organophosphorus components as ethion and diazinon.<br />

Subsequently chlorfenvinphos, fensulfothion, carb<strong>of</strong>uran, fon<strong>of</strong>os, and chlorpyrifos<br />

have been used in the seed furrow; and parathion, diazinon, naled, or malathion have<br />

been used as sprays against the adults <strong>of</strong> the second and third generation. Resistance<br />

to several <strong>of</strong> these insecticides has developed and, while the level <strong>of</strong> resistance has not<br />

been as great as with the cyclodienes, commercial control is becoming increasingly<br />

difficult (Harris & Svec 1976).<br />

The use <strong>of</strong> the sterile-male approach for insect control was first successfully applied<br />

(Knipling 1960) against a livestock pest and since then its use has been explored for<br />

many insect species. In the case <strong>of</strong> the onion maggot, McClanahan & Simmons<br />

(1966) and Noordink (1966) showed that the insect could be sterilized without serious<br />

effects on the longevity <strong>of</strong> the adult. In Holland, Noordink and his associates<br />

continued studies on the sterile-male method for control, and by 1980 had completed<br />

laboratory and field testing to the point where a commercial venture in marketing the<br />

programme to the growers is now in place (Noordink 1971, 1974; Ticheler el al. 1974;<br />

Loosjes 1976).<br />

In <strong>Canada</strong>, research on sterile-male control for the onion maggot began as a<br />

cooperative venture between Agriculture <strong>Canada</strong> (C.R. Harris) and the University <strong>of</strong><br />

Guelph (F.L. McEwen) in 1971. In the initial stages <strong>of</strong> the study, mass rearing<br />

procedures were perfected (Harris & Svec 1976), and laboratory tests determined that<br />

when pupae reared under diapausing conditions were irradiated (Cobalt 60) after 6<br />

days post-diapause (24°C) at 2 or 4 krad, the males were sterilized and the females<br />

made unreproductive. Tests established that males produced from these pupae were<br />

competitive and lived about as long as those emerging from non-irradiated pupae.<br />

Initial field releases <strong>of</strong> sterilized pupae were begun in 1973 and, using fluorescent<br />

dyes as markers on released flies, the field distribution pattern was also determined.<br />

The survival <strong>of</strong> released flies was monitored and eggs collected in the field and hatch<br />

determined. The initial results indicated that the flies did not disperse widely (most<br />

recaptures were within 75 m <strong>of</strong> release site) and that the hatchability <strong>of</strong> eggs in the<br />

release area was reduced. In addition, an economic study was undertaken that showed<br />

that if the sterile-male programme was effective, it was an economically sound<br />

approach (King 1973).<br />

Field tests were expanded with 6000000 pupae released in 1974 and 10 000 000 in<br />

1975. While it was clear that egg hatch was reduced, bird predation on released pupae<br />

and immigration <strong>of</strong> wild flies into the release area resulted in inconsistent results.<br />

Extensive monitoring indicated, however, that when the ratio <strong>of</strong> wild to released flies<br />

approached 1: I, the egg hatch was reduced near 50%.<br />

29


30 F. L. McEwen<br />

Literature Cited<br />

Since rearing facilities for continuing field releases at the 1974 and 1975 levels were<br />

no longer available, the release portion <strong>of</strong> the programme was curtailed in subsequent<br />

years and efforts concentrated on learning more about the field ecology <strong>of</strong> wild and<br />

released flies, studying mating habits and determining a release procedure in which<br />

bird predation would not be such an important negative factor on the programme.<br />

These studies have demonstrated that the pupae can be released by air or as adults by<br />

air and, hopefully, will not be so subject to predation as ground release <strong>of</strong> pupae in<br />

designated sites. It has been shown also that the female fly mates only once (Martin<br />

1981) and that in the field, much <strong>of</strong> the survival <strong>of</strong> third generation larvae to<br />

overwintering pupae results from cull and volunteer onions left in the field after<br />

harvest. Thus fall sanitation is an essential component in a programme to reduce onion<br />

maggot populations.<br />

Future plans are to release approximately 100 000 000 sterilized flies in an isolated<br />

onion-growing area containing about 300 ha <strong>of</strong> onions in each <strong>of</strong> the years 1982 and<br />

1983. To accommodate this, a new biological control laboratory has been built at the<br />

University <strong>of</strong> Guelph with major facilities and space for mass rearing the onion<br />

maggot. These releases will be closely monitored and serve as a demonstration project<br />

to determine the feasibility <strong>of</strong> the sterile-male method <strong>of</strong> control. If successful on the<br />

limited area chosen, the programme will be expanded to the major onion areas in<br />

Ontario.<br />

Finlayson, D.G. (1959) Chemical control <strong>of</strong> the onion maggot in onions grown from seed in various types <strong>of</strong> soil in Northwestern America in<br />

1955-56. Journal <strong>of</strong> Economic Entomology 52, 851-856.<br />

Harris, C.R.; Svec, H. (1976) Onion maggot resistance to insecticides. Journal <strong>of</strong> Economic Entomology 69, 617-620.<br />

King, D.L.J. (1973) An economic appraisal <strong>of</strong> the sterile-male technique <strong>of</strong> onion maggot control on the Holland Marsh <strong>of</strong> Ontario. M.Sc.<br />

Thesis, University <strong>of</strong> Guelph, 176 pp.<br />

Knipling, E.F. (1960) The eradication <strong>of</strong> the screw-worm Oy. Scientific American 203, 54-61.<br />

Loosjes, M. (1976) Ecology and genetic control <strong>of</strong> the onion Oy Delia antiqua Meigen. Wageningen, PUDOC; 94 pp.<br />

Martin, J.S. (1981) Mating beha\;our <strong>of</strong> the onion maggot Oy Hylemya antiqua (Meigen) in laboratory conditions. M.Sc. Thesis, University<br />

<strong>of</strong> Guelph, 43 pp.<br />

McCIanaham, R.J.; Simmons, H.S.(1966) Sterilization <strong>of</strong> onion maggots by irradiation with cesium-137. Canadian Entomologist 98, 931-935.<br />

Noordink, J.Ph.W. (1971) Irradiation, competitiveness, and the use <strong>of</strong> radioisotopes in sterile-male studies with the onion Oy, Hylemya<br />

antiqua M. In: Sterility principle forinsect control or eradication. International Aromic Energy:Agency,<br />

232-328.<br />

Noordink, J.Ph.W. (1974) Onderzoek met behulp van radiactivc isotopen. Jaarversl. 1973. Mededelingen Instituut voor Plantenziektenkundig<br />

Onderzoek, Wageningen, 119-121.<br />

Ticheler, J.; Loosjes, M.; Noordink, J.Ph. W.; Noorlandcr, J.; Theunissen, J. (1974) Field experiments with the release <strong>of</strong> sterilized onion<br />

Dies Hylemya antiqua (Meig.). In: The sterile-insect technique and its field <strong>of</strong> applications International<br />

Atomic Energy Agency, 103-107.


Pest Status<br />

Background<br />

Chapter 10<br />

Entomoscelis americana Brown, Red<br />

Turnip Beetle (Coleoptera:<br />

Chrysomelidae)<br />

G.H. GERBER<br />

The red turnip beetle, Entomosce/is americana Brown, is an oligophagous insect<br />

which feeds on Cruciferae and is an occasional pest <strong>of</strong> rape, Brassica campestris L. and<br />

B. napus L., mustard, B. juncea (L.) Czern. and B. Itirta Moench, and cruciferous<br />

garden crops in western <strong>Canada</strong> (Stewart 1973, Gerber & Obad<strong>of</strong>in 1981a, 1981b). The<br />

larvae and adults feed on seedling plants in May and June before the adults aestivate in<br />

July, and the adults feed on mature plants in August and September after aestivation.<br />

The only damage <strong>of</strong> economic significance is caused by the adults in June (Gerber 1974,<br />

1976).<br />

The red turnip beetle is native to North America and in <strong>Canada</strong> occurs south <strong>of</strong> the<br />

Arctic climatic zone and west <strong>of</strong> longitude 96°W (Stewart 1973). The beetle overwinters in<br />

the egg stage on the surface <strong>of</strong> the soil. The eggs hatch in late April and early May,<br />

shortly after the snow has melted, but usually before rape, mustard, and cruciferous<br />

garden crops are planted (Gerber & Obad<strong>of</strong>in 1981a). Larval development normally is<br />

completed by the end <strong>of</strong> May. The larvae feed on volunteer rape and mustard<br />

seedlings, and on cruciferous weeds in fields that contained these plants the previous<br />

year (Tumock et al. 1979). The adults emerge during the first 3 weeks <strong>of</strong> June and feed<br />

for 2-3 weeks, usually in the same fields in which the larvae fed. If the supply <strong>of</strong> food<br />

plants is not adequate. the adults will move into nearby new fields containing their host<br />

plants. At the end <strong>of</strong> June, the adults enter the soil to aestivate for about a month. They<br />

reappear in late July and August; disperse to new fields <strong>of</strong> rape, mustard, or garden<br />

Cruciferae; and mate and lay eggs until late October.<br />

The only known natural enemies <strong>of</strong> E. americana attack the adult stage: a pathogen<br />

(microsporidia (Gerber 1975» and a predator (adults <strong>of</strong> the carabid beetle, Pterosticltus<br />

adstrictus (W.J. Tumock, 1975. personal communication». These two natural enemies<br />

have been encountered on only a few occasions in the field and consequently are not<br />

considered to be important control agents.<br />

Manolache (1941) reported two tachinid parasitoids in the larvae and pupae <strong>of</strong> a<br />

European species <strong>of</strong> Entomoscelis (E. adonidis Pall.) in Romania: Meigenia mllIabilis<br />

Fall. and Marcltantia clta/cltonta Meig. However, the incidence <strong>of</strong> parasitism by M.<br />

mutabilis was low «6.0%) and Manolache considered E. adonidis to be only a<br />

secondary host for this tachinid (a heteroxenous species). Manolache did not report on<br />

the abundance and importance <strong>of</strong> the other tachinid. Brovdii (1977) found up to 38% <strong>of</strong><br />

the last instar larvae <strong>of</strong> E. adonidis parasitized by tachinids in the Ukraine. but the<br />

tachinids were not identified.<br />

Studies were initiated in 1973 to identify the natural enemies <strong>of</strong> E. adonidis in<br />

Europe and to determine whether there are any parasitoids <strong>of</strong> this beetle which are<br />

suitable for introduction into <strong>Canada</strong> as control agents for E. americana. The<br />

parasitoids <strong>of</strong> E. adonidis were considered to be potential control agents for E.<br />

americana, because the life histories, phenology. and host plants <strong>of</strong> these two species<br />

31


32 G. H. Gerber<br />

Recommendations<br />

Literature Cited<br />

are similar (Manolache 1941; Stewart 1973; Brovdii 1977; Gerber & Obad<strong>of</strong>in 1981a,<br />

1981b). These studies have not been successful, because populations <strong>of</strong> E. adonidis<br />

have been extremely scarce outside <strong>of</strong> the Soviet Union since 1973. Only a small<br />

number <strong>of</strong> specimens <strong>of</strong> a species <strong>of</strong> Meigenia have been found parasitizing the larvae.<br />

Collections <strong>of</strong> E. adonidis have not yet been made in the Soviet Union.<br />

It is recommended that every attempt be made to collect additional larvae and adults <strong>of</strong><br />

E. adonidis from Europe, especially from the Soviet Union, to identify the parasitoids<br />

<strong>of</strong> this beetle. If M. mutabilis is the only parasitoid which can be collected in sufficient<br />

number for release in <strong>Canada</strong>, research is needed to determine its life history and<br />

whether it is truly heteroxenous. If M. mutabilis is heteroxenous, it would not be a<br />

suitable control agent for E. americana.<br />

Efforts also should be made to collect another species <strong>of</strong> Entomoscelis, E. suturalis,<br />

which occurs on Cruciferae in Europe. The life history <strong>of</strong> E. suturalis is similar to<br />

those for E. americana and E. adonidis (Brovdii 1977), but there is no information in<br />

the literature on its natural enemies.<br />

Brovdii, V.M. (19n) Fauna <strong>of</strong> the Ukraine - Chrysomelinae leaf beetles. Ukrainian S.S.R. Academy <strong>of</strong> Sciences. Institute <strong>of</strong> Zoology.<br />

Naukova Dumka. Kiev 19(16), 1-388. (In Ukrainian).<br />

Gerber, G.H. (1974) Red turnip beetle on rape. Canadex 149.622.<br />

Gerber, G.H. (1975) Occurrence <strong>of</strong> a microsporidian infection in the red turnip beetle. Entomo.fcelis americana (Coleoptera: Chrysomelidae).<br />

Canadian Entomologist 107, 1081-1082.<br />

Gerber. G.H. (1976) Effects <strong>of</strong> feeding by adults <strong>of</strong> the red turnip beetle, Entomosctlis americana Brown (Coleoptera: Chrysomelidae). during<br />

late July and August on the yield <strong>of</strong> rapeseed (Cruciferae). Manitoba Entomologist 10,31-35.<br />

Gerber, G.H.; Obad<strong>of</strong>in. A.A. (198Ia) Growth, development and survival <strong>of</strong> the larvae <strong>of</strong> the red turnip beetle, Entomoscelis americana<br />

(Coleoptera: Chrysomelidae) on Brassica campestris and B. napus (Cruciferae). Canadian Entomologist<br />

lB. 395-406.<br />

Gerber. G.H.; Obad<strong>of</strong>in. A.A. (198lb) The suitability <strong>of</strong> nine species <strong>of</strong> Cruciferae as hosts for the larvae <strong>of</strong> the red turnip beetle. Entomosctlis<br />

americana. (Coleoptera: Chrysomelidae). Canadian Entomologist 113.407-413.<br />

Manolache. F. (1941) Research on the morphology. biology. and control <strong>of</strong> the insect, Entomoscelis adonidis Pall .• in Romania. Methods.<br />

Guidance and Investigations. Institute <strong>of</strong> Agronomic Research Romania (Bucharest) 71. 1-205. (In<br />

Romanian).<br />

Stewan. D.B. (1973) The red turnip beetle. Entomoscelis americana Brown (Coleoptera: Chrysomelidae). biology and plant relationships.<br />

M.Sc.Thesis, University <strong>of</strong> Albena. Edmonton, 86 pp.<br />

Tumock. W.J.; Gerber, G. H.; Bickis, M.; Bennett. R.B. (1979) The applicability <strong>of</strong> X-ray energy-dispersive spectroscopy to the identification<br />

<strong>of</strong> populations <strong>of</strong> red turnip beetle. Entomoscelis americana (Coleoptera: Chrysomelidae). Canadian Ento·<br />

mologist 111. 113-125.


Pest Status<br />

Background<br />

Chapter 11<br />

Euxoa messoria (Harris), Darksided<br />

Cutworm (Lepidoptera: Noctuidae)<br />

H.H. CHENG<br />

The darksided cutwonn, Euxoa messoria (Harris), is the most destructive insect pest<br />

<strong>of</strong> tobacco in Ontario and indeed throughout most tobacco-growing areas <strong>of</strong> <strong>Canada</strong>.<br />

It is a northern species which has also done great damage to a variety <strong>of</strong> crops. The<br />

principal injury is caused by larvae which attack the newly transplanted tobacco<br />

seedlings and damage the leaves or growing points and sometimes notch or sever the<br />

plants. If left uncontrolled, a low to moderate population level <strong>of</strong> this pest has resulted<br />

in 20-48% damaged plants and 14-25% loss <strong>of</strong> yield (Cheng 1971a, 1971b, 1973c,<br />

1975, 1980a). In the tobacco-growing areas, large populations <strong>of</strong> this pest are<br />

produced in the rye rotation crop, but invasion by migrating larvae is always a problem<br />

(Bucher & Cheng 1970).<br />

The bionomics <strong>of</strong> the darksided cutworm in Ontario has been published (Cheng<br />

1973a). It has one generation a year and overwinters as an egg in the soil from 6 to 12 mm<br />

below the surface. Depending upon the weather, overwintered eggs begin to hatch<br />

from the end <strong>of</strong> March to early April. Larvae in the early instars feed on the rye<br />

rotation crop, later they attack newly transplanted tobacco seedlings. The duration <strong>of</strong><br />

the larval stage is about 3 months. The mature larvae cease feeding and transform to<br />

pupae in the soil from late July to mid-August. The pupal stage lasts from 18 to 28 days.<br />

The adults emerge from mid-August to October and lay their eggs in the soil during the<br />

same period. These eggs are completely developed before winter and hatch in the<br />

following spring.<br />

Research on darksided cutwonn control in <strong>Canada</strong> in the past 10 years has centered<br />

in two areas: (1) selection and evaluation <strong>of</strong> the more effective and safe chemical<br />

insecticides (Cheng 1979, 1980a, 1980b), and (2) search and evaluation <strong>of</strong> effective<br />

biological control agents (Bucher 1970, 1971; Cheng 1973b, 1977). These approaches<br />

are reviewed in this chapter.<br />

As more information becomes available on the hazards involved in the agricultural use<br />

<strong>of</strong> insecticides, there is a greater demand for alternative methods for controlling<br />

insect pests. Consequently, laboratory and field tests with virus pathogens and<br />

Bacillus rhuringiensis preparations were conducted from 1968 to 1972 at the Research<br />

Station, Delhi, Ontario (Bucher 1970, 1971, Cheng 1973b). The results indicated that<br />

these pathogens were significantly less effective than chemical insecticides against E.<br />

messoria larvae in the field (Cheng 1973b).<br />

Studies on the biology and collection <strong>of</strong> parasitoids, predators, and diseases <strong>of</strong> E.<br />

messoria have also been conducted by Agriculture <strong>Canada</strong> at the Delhi Research<br />

Station. The degree <strong>of</strong> the influence <strong>of</strong> various biotic factors on E. messoria<br />

populations and the importance and possible roles <strong>of</strong> the natural enemies in the<br />

tobacco crop and pest relationship have been discussed (Bucher & Cheng 1971,<br />

Cheng 1973a, 1977). In general. larval mortality in E. messoria from natural enemies<br />

was low.<br />

Tobacco is the number one cash crop in Ontario. Because <strong>of</strong> the high value <strong>of</strong> the<br />

crop and the insufficient pressure <strong>of</strong> the various biotic factors on the dark sided<br />

33


34 H. H. Cheng<br />

Parasitoids<br />

Predators<br />

cutworm, the use <strong>of</strong> chemical insecticides for control <strong>of</strong> this pest is vital to tobacco<br />

production. Persistent chemical insecticides, such as DDT and endosulfan (Thiodan®,<br />

Niagara Otemicals) cannot be used on tobacco because <strong>of</strong> the danger <strong>of</strong> environmental<br />

contamination and undesirable residues in tobacco (Cheng & Braun 1977, Frank et al.<br />

1977). Therefore, research in the evaluation <strong>of</strong> insecticidal control <strong>of</strong> cutworms in<br />

tobacco has been directed towards the more effective and safe chemical insecticides<br />

(Cheng 1973c, 1975, 1979, 1980a, 1980b). For example, the broad-spectrum pyrethroids,<br />

permethrin, and cypermethrin are very toxic to the cutworm larvae and nonhazardous,<br />

and have been registered and recommended for both preplanting (on the<br />

rye cover crop or on soil surface) and postplanting (on tobacco plants) for control <strong>of</strong><br />

cutworms.<br />

From 1973 to 1977, 14 species <strong>of</strong> insect parasitoids were reared from E. messoria<br />

larvae or pupae in Ontario; they all were primary and internal parasitoids, namely,<br />

Apanteles laeviceps Ashmead, A. militaris Walsh, Meteorus communis (Cresson) and<br />

M. levil'entris (Wesmael) (Hymenoptera: Braconidae); Arenetra rufipes vernalis<br />

Walley, Campoletis flavicinctus (Ashmead), Campoletis sp., Enicospilus sp., and<br />

Eutanyacra sllturalis (Say) (Hymenoptera: Ichneumonidae); Copidosoma bakeri<br />

(Howard) (Hymenoptera: Encyrtidae); Muscina stabulans (Fall.) (Diptera: Muscidae);<br />

Linnaemya compta (Fall.), Winthemia rll/opicta (Bigot), and W. deilephilae (O.S.)<br />

(Diptera: Tachinidae). Of these 14 species, Arenetra rufipes vernalis, Copidosoma<br />

bakeri, and Enicospilus sp. were the most common species reared from E. messoria<br />

larvae and were present in considerable numbers every year. The total parasitism by<br />

these three species was about 22% (Cheng 1977) and were considered to be the most<br />

important parasitoids <strong>of</strong> E. messoria.<br />

From the economic point <strong>of</strong> view, A. rllfipes vernalis is considered to be the most<br />

effective parasitoid <strong>of</strong> the darksided cutworm in the current generation in Ontario. It<br />

kills the sixth-instar larvae and essentially prevents the cutworm from causing the<br />

maximum damage to the tobacco crop, because the seventh instar <strong>of</strong> this pest does<br />

more damage to the tobacco crop than all the other instars combined. Enicospi/us sp.<br />

kills its host in the seventh instar or prepupal stage, and it reduces some losses caused<br />

by the cutworm. Conversely, C. bakeri increases the crop losses by the current<br />

generation, because the cutworms attacked by C. bakeri consumed 27.5% more food<br />

than did normal cutworms (McMillan 1930); however, it will reduce the population<br />

levels <strong>of</strong> the following generation <strong>of</strong> this pest. The rest <strong>of</strong> the parasitoids were present<br />

in small numbers or reared only once during the period <strong>of</strong> this study and it is likely that<br />

they are not important.<br />

Several birds, insects, and spiders were observed to attack the larvae and adults <strong>of</strong> the<br />

dark sided cutworm (Cheng 1973a). Birds periodically landed in the fields in search <strong>of</strong><br />

larvae in the early morning; most frequently recorded were starlings and grackles. Two<br />

species <strong>of</strong> ground beetles, Calosoma calidum (F.) and Harpalus caliginosus F.<br />

(Coleoptera: Carabidae), were the most important insect predators. Both the adults<br />

and larvae <strong>of</strong> the ground beetles destroyed a large number <strong>of</strong> cutworms. Moths <strong>of</strong> the<br />

darksided cutworm were preyed upon by spiders, which were identified as Leiobunum<br />

vittatum (Say), Phalangium opilio (L.), Hadrobllnus maculostlS (Wood), and Odie/lus<br />

pictus (Wood), as the moths rested in their hiding places. However, definitive studies<br />

on the effectiveness <strong>of</strong> these predator complexes in the regulation <strong>of</strong> E. messoria<br />

populations are lacking. Nevertheless, they are assumed to be valuable biological<br />

control factors.


Diseases<br />

Microbial Insecticide<br />

EIlXO(l messoria (Harris). 35<br />

Several insect pathogens have been isolated from the diseased larvae <strong>of</strong> E. messoria<br />

collected in the fields at Delhi, Ontario (Bucher & Cheng 1971). The bacterial diseases<br />

are common and the bacteria most commonly isolated from infested larvae are Bacillus<br />

cereus Frankland & Frankland, Enterobacter cloacae (Jordan), Enterobacter aerogenes<br />

(Kruse), Klebsiella pneumoniae (Schroeter), Streptococcus faecalis (Andrewes<br />

& Horder), Pseudomonas j1uorescens Migula, Pseudomonas spp., Bacillus splraericus<br />

Neide, and Achromobacter spp. The infected haemocoels <strong>of</strong> cutworm larvae usually<br />

contained one or a mixture <strong>of</strong> two or more bacterial species. B. cereus is a well-known<br />

pathogen <strong>of</strong> many insects (Heimpel & Angus 1963); E. aerogenes and P. fluorescens<br />

have been considered as potential pathogens <strong>of</strong> a number <strong>of</strong> insects (Bucher 1963); the<br />

pathogenicity <strong>of</strong> the other species is very questionable and they probably occurred by<br />

chance in the haemocoel <strong>of</strong> sick larvae <strong>of</strong> E. messoria. The microsporidial disease,<br />

Nosema sp., is less common; the fungus disease, Sorosporella uvella (Krass.), is rare;<br />

and the virus diseases are virtually absent.<br />

Although larvae <strong>of</strong> E. messoria collected in the fields in Ontario are free from virus<br />

diseases, they are susceptible to two virus diseases, nuclear polyhedrosis and<br />

granulosis virus, originally isolated from a related cutworm, E. ochrogaster (Guenee)<br />

(Bucher 1970). In cooperation with Dr. G.E. Bucher <strong>of</strong> Agriculture <strong>Canada</strong> Research<br />

Institute, Belleville, Ontario, field tests using preparations <strong>of</strong> nuclear polyhedrosis<br />

virus and granulosis virus were conducted in 1969 and 1970 at the Research Station,<br />

Delhi, Ontario. These tests showed that both viruses would protect tobacco seedlings<br />

from severe cutworm damage when applied on the green rye in high concentrations,<br />

but they were less effective than chlorpyrifos (Dursban® or Lorsban l8l , Dow Chemical<br />

<strong>of</strong> <strong>Canada</strong> Ltd.) treatment and better than the untreated control.<br />

In 1971 and 1972, attempts were made to introduce both virus diseases into healthy<br />

populations <strong>of</strong> E. messoria by spraying the virus suspensions on tobacco trap-plants<br />

growing in a rye field. Consequently, a considerable number <strong>of</strong> E. messoria larvae<br />

collected with tobacco trap-plants from areas where viruses were introduced in 1971<br />

and 1972 were infested with virus. This indicated that a large amount <strong>of</strong> virus survival<br />

occurred and that virus introduction is feasible. Unfortunately, this programme was<br />

discontinued after the Research Institute, Belleville, Ontario, closed in the fall <strong>of</strong> 1972.<br />

In 1969, laboratory experiments were conducted to determine the susceptibility <strong>of</strong> the<br />

various larval stages <strong>of</strong> E. messoria in Ontario to four commercial preparations <strong>of</strong><br />

Bacillus thuringiensis Berliner: Thuricide® 9OTS, Thuricide® HP, Biotrol l8l BTB 183,<br />

and Dipel® (Sandoz <strong>Canada</strong>). In these experiments, first-instar to third-instar larvae<br />

that fed on rye leaves treated with all four B. thuringiensis preparations were found to<br />

be susceptible at all rates applied (Cheng 1973b). Mortality <strong>of</strong> fourth-instar to seventhinstar<br />

larvae fed treated tobacco leaves was low.<br />

From the results <strong>of</strong> laboratory experiments, it was apparent that preparations <strong>of</strong> B.<br />

thurittgiensis showed potential as a possible alternative to chemical insecticides for the<br />

control <strong>of</strong> the early instar larvae <strong>of</strong> E. messoria. Therefore, a field test was conducted<br />

in 1970 at the Agriculture <strong>Canada</strong> Research Station, Delhi, Ontario, to determine<br />

whether these preparations could be as effective as chlorpyrifos for control <strong>of</strong> this pest<br />

when applied on the rye cover crop in spring. Data showed that B. tlruringiensis<br />

preparations as applied in the field for control <strong>of</strong> E. messoria larvae were relatively<br />

ineffective as compared with chlorpyrifos (Cheng 1973b). The failure <strong>of</strong> these<br />

preparations in the field trial is probably attributable to the different environmental<br />

conditions and the much more complex ecosystem in the field than in the laboratory.


Blank Page<br />

38


Pest Status<br />

Background<br />

Releases and Recoveries<br />

Evaluation <strong>of</strong> Control Attempts<br />

Recommendations<br />

Chapter 12<br />

Forficula auricularia L. , European<br />

Earwig (Dermaptera: Forficulidae)<br />

R.F. MORRIS<br />

This report is an update on the European earwig, Forficula auricularia L., and its<br />

parasitoid, Bigonichela selipennis (Fallen), in Newfoundland as last reported by<br />

Morris (1971). This pest continued to spread throughout Newfoundland, and is now<br />

found in the following communities: St. John's, Wedgewood Park, Mt. Pearl, Kilbride,<br />

Petty Harbour, Logy Bay, Topsail, Long Pond, Torbay, Renews, Bay Roberts,<br />

Clarkes Beach, Grand Bank, Fortune, Grand Falls, Stephenville.<br />

The European earwig was first discovered in Newfoundland at St. John's in 1948. The<br />

parasitoid, Bigonichela selipennis, was introduced from British Columbia to Newfoundland<br />

in 1951, 1952, and 1953. It was successfully established but parasitism was<br />

extremely low during the period 1955-59 (Morris 1971). Because <strong>of</strong> a slow build-up,<br />

supposedly hardier strains <strong>of</strong> the parasitoid were obtained by the Research Institute at<br />

Belleville from Sweden and Switzerland and released at St. John's in 1959, 1961, and<br />

1963.<br />

The use <strong>of</strong> traps and attractants to complement earwig control with parasitoids were<br />

investigated and reported on by Morris (1965).<br />

No additional releases <strong>of</strong> Bigollichela selipennis have been made since those recorded<br />

by Morris (1971). Studies on its establishment and population increase were continued<br />

during the period from 1969 to 1978. Results are shown in Table 5.<br />

The introduced parasitoid, Bigonichela selipellllis, has successfully established itself<br />

with a maximum parasitism <strong>of</strong> 16.4% in 1973. Although the initial parasitism was low,<br />

4.4% in 1969, a steady increase was observed until 1973. There has been a considerable<br />

reduction in earwig populations in the study area since 1972, making it impossible to<br />

collect as large samples (4000 earwigs) as were used in the earlier years, and this<br />

reduction is no doubt due to the high levels <strong>of</strong> parasitism during the 1972-76 period.<br />

There has been a continuing spread <strong>of</strong> earwigs from the original site in St. John's with<br />

populations now established 420 km and 775 km from St. John's. Studies should be<br />

initiated to determine if the parasitoid has been introduced with its host to Grand<br />

Bank, Fortune, Grand Falls, and Stephenville. If it has not been introduced naturally,<br />

steps should be taken to artificially introduce Bigonichela selipellnis to these areas.<br />

39


Pest Status<br />

Background<br />

Releases and Recoveries<br />

Parasitoids<br />

Pathogens<br />

Chapter 13<br />

Hypera postica (Gyllenhal), Alfalfa<br />

Weevil (Coleoptera: Curculionidae)<br />

D.G. HARCOURT and J.e. GUPPY<br />

Since its invasion <strong>of</strong> the mid-Atlantic States during the early 1950s, the eastern strain<br />

<strong>of</strong> the alfalfa weevil, Hypera pos/iea (Oyll.), has become the most important pest <strong>of</strong><br />

alfalfa in eastern North America. Of Eurasian origin, this biotype spread rapidly<br />

northward and in little more than a decade reached Ontario and Quebec where it began<br />

to alarm dairy farmers in the lower Great Lakes region. From 1967 to 1974,<br />

populations spread throughout southern Ontario, increasing at geometric rates.<br />

However, by the mid-seventies. this pattern <strong>of</strong> unregulated growth had been checked<br />

by the combined action <strong>of</strong> natural enemies, and during the past 5 years numbers <strong>of</strong> the<br />

pest have been constrained at a density level that fluctuates about the economic<br />

damage threshold. As a result. outbreaks have been restricted. However, scattered<br />

epicentres <strong>of</strong> infestations still exist and sudden population flare-ups are prone to occur.<br />

In 1957. the Beneficial Insects Research Laboratory. Northeastern region,<br />

U.S.D.A .• began to make releases <strong>of</strong> exotic parasitoids against the weevil. A total <strong>of</strong><br />

13 species was released, several <strong>of</strong> which became established (Abu & Ellis 1976).<br />

Discovery <strong>of</strong> the pest in Ontario in 1967 prompted recolonization here, <strong>of</strong> those species<br />

that had become established in the northeastern United States (Williamson 1971, 1972).<br />

Six species <strong>of</strong> parasitoids were released in various localities in Ontario during 1970 and<br />

1971 (Williamson 1971, 1972): Bathyplectes anurus (Thomson), B. eureulionis<br />

(Thomson), B. slenosligma (Thomson), Mieroelonus aelhiopoides Loan, M. coles;<br />

Drea, and Telras/jehus ineertus (Ratzeburg). Recoveries in 1971 included all species<br />

except M. eolesi which was not found until 1978. B. slenostigma was not found in<br />

follow-up years but all other species became firmly established. Although B. anurus<br />

was recovered the first year after its release. it has been slow to establish and spread,<br />

and its incidence <strong>of</strong> parasitism is <strong>of</strong> little importance.<br />

The egg parasitoid, Palasson luna (Girault) was not released in Ontario but has<br />

spread across the province with its host.<br />

In 1973. an infectious disease caused by the fungus Entomophthora phytonomi Arthur<br />

(Harcourt et al. 1974) was discovered in the Bay <strong>of</strong> Quinte area <strong>of</strong> southern Ontario<br />

where it caused a spectacular epizootic in populations <strong>of</strong> the weevil. New to the insect<br />

in North America. it was soon found throughout Ontario and in adjacent parts <strong>of</strong> the<br />

United States (Muka & Oyrisco 1976).<br />

41


Hypem poslica (GyllcnhaJ). 43<br />

I-Iarcourt. D.G.; Guppy. J.C.; Macleod. D.M.; Tyrrell. D. (1974) The fungus Emomopill/wrtl pi,ylollomi pathogenic to the alfalfa weevil.<br />

Hypera postim. Catracliatr Emomologisl 106. 1295-1300.<br />

Harcourt. D. G.; Guppy, J.c.; Binns, M. R. (1977) TIle analysis <strong>of</strong> inlrageneration change in eastern Ontario populations <strong>of</strong>lhc alfaUa weevil.<br />

Hypera postiCCl (Coleoptera: Curculionidae). Canadian Elllomologist 109, 1521-1534.<br />

Harcourt. D.G.; Ellis, C.R.; Guppy. J.C. (1980) Distribution <strong>of</strong> Microclonlls aelltiopoitles.1I parusitoid <strong>of</strong> the alfalfa weevil (Coleoptera:<br />

Curculionidlle) in Ontario. Proceedings <strong>of</strong> tlte Emomologicltl <strong>Society</strong> <strong>of</strong> Omario (1979) lID, 35-39.<br />

loan. C.C. (1971) SitOIlCl cylindricollis Fahr., sweelclover weevil, and Hypera postica (Gyll.) •• lIf.llfll weevil (Coleoptera: Curculionidae). In:<br />

Biological conlrol progrummcs against insects and weeds in Cllnadll 1959-68. Commonwealtll blStitl4le<br />

<strong>of</strong> Biological COlllml Technical Commwlicatiotl 4. 43-46.<br />

Muka. A.A.; Gyrisco, G.G. (1976) Report presenled III 13th Northeast Regional Alfalfa Insects Conference, Newark, Delaware.<br />

Williamson, G.D. (1971) Insect liberalions in <strong>Canada</strong>, 1970. Parasites and predators. Canacia AgricllllIlre Liberatioll Bulietin 34. 16 pp.<br />

Williamson. G.D. (1972) Insect liberations in <strong>Canada</strong>. 1971. Parasites and predators. Canacia Agrialllllre LiberatiOlJ Bulietin 35. 16 pp.


Blank Page<br />

44


Pest Status<br />

Background<br />

Chapter 14<br />

Lygus spp.,<br />

Miridae)<br />

C.H. CRAIG and c.c. LOAN<br />

Plant Bugs (Heteroptera:<br />

About 30 species <strong>of</strong> Lygus are known to occur from coast to coast in <strong>Canada</strong> (Kelton<br />

1975). Most <strong>of</strong> the species feed on a variety <strong>of</strong> herbaceous and woody plants. some are<br />

host specific or are limited to related plant species. Many species <strong>of</strong> Lyglls are<br />

economically important as pests <strong>of</strong> fruits. vegetables. and tobacco crops, and <strong>of</strong> legume<br />

crops. primarily those grown for seed. Of the dozen species found on alfalfa the most<br />

abundant and important as pests are L. borealis Kelton and L. lineo/aris (Palisot de<br />

Beauvois), with L. unctuosus (Kelton). L. desertinus Knight, and L. shulli Knight <strong>of</strong><br />

lesser importance.<br />

Lygus overwinter as adults in the shelter <strong>of</strong> plant litter on the soil surface. In spring<br />

they feed on the early plants and move to alfalfa when it becomes suitable for feeding<br />

and oviposition.<br />

In <strong>Canada</strong> the number <strong>of</strong> generations <strong>of</strong> Lygus per year is not well documented<br />

because nymphal taxonomy is not known. It is generally conceded. however. that the<br />

species are univoltine in northern latitudes and bivoltine. or largely so. south <strong>of</strong> about<br />

latitude 51°N.<br />

Lygus feeding injures the vegetative and reproductive parts <strong>of</strong> the plant causing<br />

stunting. bud-blast and premature nower drop. and reduced yield and quality <strong>of</strong> seed.<br />

In most years and in most alfalfa seed fields. Lygus infestations exceed the economic<br />

threshold and control measures are essential.<br />

In Poland. Lygus rugulipennis Poppius. the dominant species on alfalfa. is parasitized<br />

by the braconids PeristemlS digoneUlis Loan. P. stygicus Loan. and P. rubricol/is<br />

(Thomson). The former two are bivoltine and attack both generations <strong>of</strong> host; the<br />

latter is univoltine and attacks only the first generation host (Bilewicz-Pawinska<br />

1969. 1975).<br />

In <strong>Canada</strong>, at Belleville. Ontario. L. lineo/aris. in the first generation on forage<br />

legumes. is parasitized by Peristenus pal/ipes (Curtis) and in the second generation on<br />

several weed species. by P. pseudopal/ipes (Loan) (Loan 1965. 1970).<br />

In 1975 in a survey <strong>of</strong> first-generation Lygus populations in alfalfa seed fields in<br />

Saskatchewan and Alberta. P. pal/ipes was the only euphorine parasitoid recovered.<br />

Incidence ranged from 3% to 49%. with an average <strong>of</strong> 21 % in the southern shortgrass<br />

prairie region. 28% in the central mixed prairie region, and 11% in the northern<br />

parkland-mixed forest region (Loan & Craig 1976). Surveys since that time have noted<br />

the occurrence <strong>of</strong> euphorine parasitism probably by P. pallipes, in second-generation<br />

Lygus in the southern prairie region (Craig. 1980. unpublished data).<br />

45


46 C. H. Craig and C. C. Loan<br />

Releases and Recoveries<br />

Recommendations<br />

Table 6<br />

A programme to supplement the natural parasitism in Lygus populations in western<br />

<strong>Canada</strong> was started in 1977 with the first collection and shipment to Ottawa <strong>of</strong> parasitoids<br />

<strong>of</strong> European Lyglls, and in 1978 with the first liberations <strong>of</strong> these in western<br />

<strong>Canada</strong>. Table 6 shows the progression <strong>of</strong> liberations in the period 1978-81 inclusive.<br />

Liberations <strong>of</strong> Peristenus spp. in alfalfa against Lygus spp. in western<br />

<strong>Canada</strong>.<br />

Date <strong>of</strong> liberation Location <strong>of</strong> liberation site<br />

7 June 1978<br />

9 June 1978<br />

13 June 1978<br />

21 June 1978<br />

14 June 1979<br />

21 June 1979<br />

29 June 1979<br />

29 May 1980<br />

3 June 1980<br />

10 June 1980<br />

10 June 1980<br />

6 June 1981<br />

6 June 1981<br />

!<br />

}<br />

Moose Jaw, Saskatchewan; 50 o N;<br />

shortgrass prairie; dry<br />

Saskatoon, Saskatchewan; 52°N;<br />

mixed prairie; dry<br />

Tilley, Alberta; 50 0 30'N;<br />

shortgrass prairie;<br />

irrigated<br />

Shellbrook, Saskatchewan; 53°N;<br />

parkland-mixed forest; dry<br />

Ardath, Saskatchewan; 51°30'N;<br />

mixed prairie; dry<br />

Shellbrook, Saskatchewan<br />

(supplement to 1979 release)<br />

Saskatoon, Saskatchewan<br />

mixed prairie; irrigated<br />

Saskatoon, Saskatchewan<br />

(supplement to 1978 release)<br />

Yellow Creek, Saskatchewan;<br />

53°N parkland; dry<br />

Saskatoon, Saskatchewan<br />

(supplement to 1980 release)<br />

Species<br />

P. digoneLltis<br />

P. digoneLltis<br />

P. digoneutis<br />

P. stygicus<br />

P. digoneLltis<br />

P. srygicus<br />

P. digoneLltis<br />

P. digoneutis<br />

P. digoneutis<br />

P. digoneutis<br />

P. digoneutis<br />

P. digoneutis<br />

Number<br />

The imported parasitoids were released into Lygus populations breeding in alfalfa<br />

fields used for seed production. The lifespan <strong>of</strong> the sites for sampling purposes is about<br />

5 years, and during that time they are subject to normal alfalfa seed cropping practices.<br />

Intensive sampling for presence <strong>of</strong> adult Peristenus and mass collection <strong>of</strong> host<br />

nymphs for parasitoid rearing was started the year follOwing the initial liberations and<br />

continued for at least three ensuing years. In 1979-80, a total <strong>of</strong> 274 parasitoid adults,<br />

either captured at the liberation sites or reared from Lygus hosts collected from alfalfa<br />

at the liberation sites, were the native species, Peristenus pallipes. No introduced<br />

parasitoids have been recovered.<br />

The introduction and liberation <strong>of</strong> Peristenus parasitoids will terminate in the spring <strong>of</strong><br />

1981. During the next few years attempts will be made to recover the introduced<br />

parasitoids and to evaluate the programme.<br />

197<br />

177<br />

541<br />

41<br />

159<br />

23<br />

167<br />

261<br />

80<br />

138<br />

125<br />

115


Literature Cited<br />

Lyglls spp., 47<br />

Bilewicz-Pawinska. T. (1969) Natural limitation <strong>of</strong> LygllS TIlgu/ipel"'is Popp. by groups <strong>of</strong> UiophTon pallipes Curtis on the rye crop fields.<br />

Ekologia /'olska, Series .. 1 17.811-815.<br />

Bilcwicz-Pawinska. T. (1975) Distribution <strong>of</strong> insect parasites l'eristellllS Foerster lind MesochoTlis Gravenhorst in Poland. Bulletin lie<br />

I'ACtldemie Polonaise des Sciences 23. 823-827.<br />

Kelton. L.A. (1975) The LygllS bugs (genus Lygus Hahn) <strong>of</strong> North America (Heteroptera: Miridac). Memoirs o/tlle Entom%gielll <strong>Society</strong> <strong>of</strong><br />

<strong>Canada</strong> 95, 101 pp.<br />

Loan. C.C. (1965) Life cycle and development <strong>of</strong> LeiopilTonpllllipes Curtis (Hymenoptera: Braconidae. Euphorinae) in five mirid hosts in the<br />

Belleville district. Proceedi/lgs <strong>of</strong> tile ElllomologiCtlI <strong>Society</strong> <strong>of</strong> Ontario (1964) 95, 115-121.<br />

Loan. C.C. (1970) The new parasites <strong>of</strong> the tarnished plant bug in Ontario: Leioplirol/1JSlmdopallipes and Euphoriantllygivora (Hymenoptera:<br />

Braconidae. Euphorinae). Prm:eel!i'lgs af the Elllolllological <strong>Society</strong> <strong>of</strong> Olllario 100, 188-195.<br />

Loan, C.C.; Craig. C.H. (1976) Euphllrine parasitism <strong>of</strong> Lyglls spp. in lllflllfil in western CllOllda (Hymenoptera: Braconidae; Heteroptem:<br />

Miridac). NtltllTll/iste Cal/aditlll 103, 497-500.


Blank Page<br />

48


Introduction<br />

Pest Status<br />

Background<br />

Chapter 15<br />

Mamestra configurata Walker, Bertha<br />

Armyworm (Lepidoptera: Noctuidae)<br />

W.J. TURNOCK<br />

The bertha armyworm, Mamestra configllrata Walker, was first recognized as a major<br />

pest <strong>of</strong> the rapeseed and canola crops, Brassica nap"s (L.) and B. campestris L., on the<br />

Canadian prairies in 1971. Little was known <strong>of</strong> the life history and natural control <strong>of</strong> this<br />

species. In 1972, research programmes were initiated at Agriculture <strong>Canada</strong> Research<br />

Stations in western <strong>Canada</strong>. Soon after, the Commonwealth Institute <strong>of</strong> Biological<br />

Control station at Delemont. Switzerland, began a study <strong>of</strong> the closely related European<br />

species, Mamestra brassicae (L.). To date, no decision has been made to introduce<br />

parasitoids <strong>of</strong> M. brassicae to <strong>Canada</strong>. This report summarizes information on the<br />

natural enemies <strong>of</strong> M. configllrata and M. brassicae and suggests future actions in<br />

biological control <strong>of</strong> the former species in <strong>Canada</strong>.<br />

M. configurata is native to North America and in <strong>Canada</strong> occurs from Manitoba to<br />

British Columbia (Wylie & Bucher 1977). Its native hosts are unknown, but it has been<br />

reported to feed on a wide variety <strong>of</strong> cultivated plants and introduced weeds (Beirne<br />

1971). On the Canadian prairies sporadic infestations occurred on various crops before<br />

1940. Since then attacks have largely been confined to fields <strong>of</strong> rapeseed and canola<br />

(Turnock & Philip 1979). From 1940 to 1970 the outbreaks were local and <strong>of</strong> short<br />

duration. From 1971 to 1974 an outbreak <strong>of</strong> unprecedented severity occurred from<br />

western Manitoba across the parkland region to Edmonton, and from Lethbridge to<br />

the Peace River District in Alberta. Insecticides were applied to about 38 000 ha<br />

(17.7% <strong>of</strong> the seeded area) in 1971, and about 31000 ha (23.3% <strong>of</strong>the seeded area) in<br />

1972. The outbreak declined in 1973 and damaging infestations were absent by 1975.<br />

Populations increased again in 1978 and 1979, and in 1980 about 49 000 ha (2.4% <strong>of</strong> the<br />

seeded area) were sprayed.<br />

The bertha armyworm is univoltine in <strong>Canada</strong> and overwinters in the pupal stage.<br />

Adults emerge from the first week <strong>of</strong> June to early August, in most years 50% <strong>of</strong> the<br />

adults have emerged by mid-July. Egg-laying, on the underside <strong>of</strong> the leaves <strong>of</strong> host<br />

plants, begins soon after emergence and the eggs hatch about 1 week later. There are<br />

six larval instars. The larvae feed mainly on the leaves, but the older instars may<br />

damage flowers and pods. Fully-fed larvae enter the soil in late August and September<br />

and pupate at depths down to 15 cm.<br />

The early reports <strong>of</strong> bertha armyworm infestations do not include lists <strong>of</strong> parasitoids or<br />

diseases (Wylie & Bucher 1977). Since 1972, collections <strong>of</strong> larvae have yielded 15<br />

species <strong>of</strong> parasitoids (Table 7). Wylie (1979) reared an additional 3 species from larvae<br />

and 5 species from pupae that had been held in field cages in an area where the bertha<br />

armyworm has never been found. Neither egg nor pupal parasitoids has been found in<br />

field collections.<br />

49


50 W. J. Turnock<br />

Table 7<br />

Table 8<br />

Parasitoids recovered from field-collected larvae <strong>of</strong> Mamestra eonfigllrata<br />

Walker in <strong>Canada</strong>.<br />

Hymenoptera<br />

Ichneumonidae<br />

Banehlls f1aveseens Cress. (M,S,A)·<br />

Braconidae<br />

Apanteles xylinlls (Say) (M)<br />

Cotesia laevicips (Ashm.) (M)<br />

Eulophidae<br />

Elllophlls nr. nebillosus (Prov.) (M)<br />

Eupleetrus bieolor Swed. (M)<br />

Diptera<br />

Tachinidae<br />

Athrycia cinerea (Coq.) (M,S,A)·<br />

Mericia ampeius (WIK.) (M,S,A)<br />

Exorista mella (Wlk.) (S.A)<br />

Phryxe peeosensis(Tnsd.) (M)<br />

P. vulgaris (Fall.) (M)<br />

Chaetogena sp. (A)<br />

C. sp. (c1aripennis (Macq.) group) (A)<br />

Lespesia sp. (A)<br />

Winthemia rufopieta (Bigot) (M)<br />

W. qlladripustulata (F.) (M)<br />

• Location <strong>of</strong> collection points in parenthesis. (M = Manitoba, S = Saskatchewan,<br />

A = Alberta).<br />

Banehus f1aveseens Cress. was the most abundant species regardless <strong>of</strong> host<br />

abundance (Table 8). This univoltine parasitoid lays its eggs in the early larval instars<br />

and completes its development before the host would normally pupate.<br />

Parasitism ('Yo) by Banehus f1aveseens Cress., Athrycia cinerea (Coq.), Mericia ampelus<br />

(Wlk.), and other species on Mamestra eonfigurata Walker in Manitoba, Saskatchewan,<br />

and Alberta, 1972-80.<br />

Parasitism ('Yo)<br />

Year(s) Province No. No. Banehus Athrycia Mericia Other<br />

Fields Larvae<br />

1972-74 Manitoba 119 1915 28.3 20.7 0.4 0.1<br />

Saskatchewan 1 59 47.5 1.7 3.4 1.7<br />

Alberta 5 506 42.5 21.7 2.4 0<br />

1975-78 Manitoba 152 100 44.0 25.0 0 2.0<br />

1979 Manitoba 43 410 6.1 9.3 0 1.7<br />

Saskatchewan • 2 1273 14.0 29.0··<br />

1980 Manitoba 51 2154 26.3 6.2 0.04 0.3<br />

Alberta 3 775 25.3 4.4 0 0<br />

Data provided by A.P. Arthur, Agriculture <strong>Canada</strong>, Research Station, Saskatoon .<br />

•• Mainly A. cinerea.<br />

Athrycia cinerea (Coq.) usually was less abundant than B. f1al'eseens when host<br />

populations were low (Table 8). but it may be as abundant in some fields during<br />

outbreaks. A. cinerea is univoltine in the Prairie Provinces. The eggs arc attached to<br />

the integument <strong>of</strong> larvae in the fourth to sixth instars. the larvae usually develop<br />

gregariously, and they mature before the host completes its feeding (Wylie 1977a).


MameSlra configurara Walker, 51<br />

Mericia ampelus (W1k.) was found in a very small percentage <strong>of</strong> bertha armyworm<br />

larvae collected during outbreaks, but has not been collected between outbreaks<br />

(Table 8). This species attaches its eggs to the leaves <strong>of</strong> plants and the larvae hatch 1-2<br />

minutes after oviposition. It appears to be bivoltine in <strong>Canada</strong>, but only the second<br />

generation parasitizes bertha armyworm. Larval survival is low, suggesting that Mamestra<br />

configurala is not a preferred host (Wylie 1977b).<br />

Pathogens have not been a major cause <strong>of</strong> mortality in the bertha armyworm. Wylie<br />

& Bucher (1977) identified fungus pathogens <strong>of</strong> the genus Entomophthora and a<br />

nuclear polyhedrosis virus as the main larval diseases in 1972. In addition, a protozoan<br />

parasite, possibly <strong>of</strong> the genus Nosema, was recovered from a few larvae.<br />

The cabbage moth, Mameslra brassicae (L.), is similar to M. configurala in its life<br />

history and phenology. M. brassicae feeds on a wide range <strong>of</strong> plants and is a major pest<br />

<strong>of</strong> cabbage in northern Eurasia. In western Europe (Switzerland, Austria, Germany)<br />

and in the warmer climatic zones <strong>of</strong> the Soviet Union (Kiev, Kharkov, Voronetz, and<br />

Krasnodar regions), it is bivoltine (CIBC Annual Project Statements 1974-80,<br />

Dzhuran 1979, Agarkov 1974, Stenin 1974, Drozda & Garnaga 1978, Tsybul'ko 1973).<br />

In cooler parts <strong>of</strong> the Soviet Union (Moscow, Gorky, Leningrad, Volgoda (Perm),<br />

Tomsk, Altai, Irkutsk, and Sakhalin regions), it is univoltine (Hori 1935, Kopvillem<br />

1960, Shchetinin 1974, Prokov'cva 1976, Kostylyova 1979). The univoltine populations<br />

lie north and the bivoltine populations lie south <strong>of</strong> the annual isotherm for 2 200" days<br />

over lOoC (Fig. 1).<br />

The parasitoid complex <strong>of</strong> M. brassicae consists <strong>of</strong> 47 species, <strong>of</strong> which 20 occur in<br />

the USSR (Razumov 1972, Kopvillem 1962). However, only 23 species are reported in<br />

the CIBC Annual Project Statements and recent literature from the Soviet Union<br />

(Table 9). Only two species, ErneSlia consobrina (Mg.) and Exelastes cinclipes (Retz.),<br />

are abundant in both areas.<br />

Ernestia consobrina occupies a similar position in the guild <strong>of</strong> parasitoids to that<br />

occupied by Mericia ampeills in North America. E. consobrina larviposits on the food<br />

plant <strong>of</strong> the host. Parasitism is most successful in third·instar larvae and M. brassicae is the<br />

major host (Kopvillem 1960, 1962, CIBC Annual Project Statement 1974); E. consobrina<br />

is the most important parasitoid <strong>of</strong> Mameslra brassicae in most <strong>of</strong> the Soviet Union<br />

(KopviJIem 1962), but not in western Europe (CIBC Annual Project Statements<br />

1974-80). It docs not occur in Primorskiy (far east) (Slabospitskii 1980). In the areas <strong>of</strong><br />

the Soviet Union where both host and parasitoid are bivoltine, it appears to have<br />

problems in synchronization with its host. In the Kicv region, Dzhuran (1979) reports a<br />

large decrcase in parasitism from the first to the second host generation while Drozda<br />

& Garnaga (1978) note an increase in the second generation. In univoltine areas, E.<br />

consobrina appears to be a more effective parasitoid than in bivoltine areas. High<br />

levels <strong>of</strong> parasitism are reported from the Leningnld, Moscow. and Volgoda (Perm)<br />

regions (Kopvillem 1960), the Gorky region (Razumov 1972), and the Irkutsk region <strong>of</strong><br />

western Siberia (Shchetinin 1974).<br />

Exelastes cinclipes occupies a similar position in the guild <strong>of</strong> parasitoids to that<br />

occupied by B. f/al'cscens in North America. E. cinclipes is bivoltine in western<br />

Europe, where it parasitizes larvac <strong>of</strong> the second generation (CIBC Annual Project<br />

Statement 1974). In the Moscow region. it is univoltine (Kopvillem 1960). This ichneumon<br />

mainly parasitizes third·instar larvae and appears to be well synchronized with its<br />

host (Kopvillem 1960). E. cinclipes is a common parasitoid in both western Europe and<br />

the Soviet Union. In westcrn Europc and areas <strong>of</strong> the Soviet Union with bivoltine hosts,<br />

it is more important than E. cOllsobrina (Tsybul'ko 1973. Agarkov 1974, CIBC Annual<br />

Project Statements 1974-78). but in cooler parts <strong>of</strong> the Soviet Union it is less abundant<br />

(Kopvillem 1962. Razumov 1972. Shchctinin 1974).<br />

Microplilis mediator (Hal.). like B. flal't'scens in North America. attacks host larvae<br />

in the early instars but unlikc it. M. mediator is multivoltine (CIBC Annual Project


Table 9<br />

Mamestra configurata Walker. 53<br />

Insect parasitoids <strong>of</strong> Mamestra brassicae Walker in western Europe and the USSR (See<br />

Fig. 1 for location <strong>of</strong> regions in the USSR).<br />

Family and Species<br />

Tachinidae<br />

Erneslia eonsobrina (Mg.)<br />

Phryxe vulgaris (Fall.)<br />

Alhrycia impressa (Wlp.)<br />

A. erYlhroeera R.D.<br />

Sip/rona erislala (F.)<br />

S. flavifrons Staeger<br />

Winlhemia quadripuslulala (F.)<br />

Exorisla larvarum (L.) (Taehina<br />

larvarum (L.»<br />

Blondelia nigripe.f (Fall.)<br />

Ceromasia sp.<br />

Ichncumonidae<br />

E.wasles cinelipes (Rctz.)<br />

Therion giganteum (Grav.)<br />

Campoplegini sp.<br />

Sleniehnellmon eulpalor (Schrank)<br />

Braconidac<br />

Microplilis mediator (Hal.)<br />

M. tubereulifer (Wcsm.)<br />

COlesia eongesla (Nees)<br />

C. glomerala (L.)<br />

MeleOru.v leviventris (Wesm.)<br />

Eulophidae<br />

ElllopllUs larvarllm L.<br />

Ellpleclrus sp.<br />

Eupleelrlls bieolor Swcd.<br />

Trichogrammatidae<br />

Triehogramma evaneseens Wcstw.<br />

Host Bivoltinc<br />

W. Europe. Krasnodar.<br />

Kiev. Saratov<br />

W. Europe<br />

Primorskiy<br />

W. Europe<br />

Krasnodar. Veronezh<br />

W. Europe<br />

W. Europe<br />

W. Europe. Kharkov<br />

Krasnodar<br />

W. Europe<br />

W. Europe<br />

Kharkov. Primorskiy<br />

Krasnodar. Voronezh<br />

Krasnodar<br />

Krasnodar. Kharkov<br />

W. Europe. Krasnodar.<br />

Voronezh. Kharkov. Kiev<br />

Region <strong>of</strong> Occurrence<br />

Host Univoltine<br />

Leningrad. Volgoda.<br />

Moscow. Gorky. Tomsk<br />

Tomsk<br />

Tomsk<br />

Tomsk<br />

Tomsk<br />

Moscow<br />

Moscow<br />

Moscow<br />

Sakhalin<br />

Moscow. Gorky. Tomsk.<br />

Sakhalin<br />

Sakhalin<br />

Gorky<br />

Sakhalin<br />

Sakhalin<br />

Gorky. Irkutsk. Altai<br />

Statement 1978). M. mediator is the most abundant parasitoid <strong>of</strong> Mamestra brassicae in<br />

Europe (CIBC Annual Project Statements 1974-80) but is not reported from the Soviet<br />

Union. It does not enter dispause to overwinter (CIBC Annual Project Statement 1980)<br />

and this characteristic may explain its absence from the Soviet Union. Microplitis tuberculifer<br />

(Wesm.) was the most abundant parasitoid in far eastern regions (Slabospitskii<br />

1980). where Mamestra brassicae is bivoltine. M. brassicae is univoltine on the island <strong>of</strong><br />

Sakhaline and is attacked by 5 species <strong>of</strong> parasitoids (Hori 1935). Only one <strong>of</strong> these. E.<br />

cinctipes. also occurs in other parts <strong>of</strong> the Soviet Union (Table 7).<br />

An egg parasitoid. TricllOgramma el'anescens Westw., occurs in western Europe<br />

and the Voronezh. Kharkov. Krasnodar, Gorky. Irkutsk. and Altai regions <strong>of</strong> the USSR<br />

(Razumov 1972. Tsybul'ko 1973. Agarkov 1974. Prokoreva 1976, CIBC Annual Project


54 W. J. Turnock<br />

Recommendations<br />

Acknowledgements<br />

Statement 1978, Kostylyova 1979). In the regions where the host is univoltine, T. evanescens<br />

generally is very rare (Prokoreva 1976). It is used as an inundative biological<br />

control agent in the USSR.<br />

The major components <strong>of</strong> the parasitoid guilds <strong>of</strong> M. configurata and M. brassicae<br />

attacking the early instar larvae include one ichneumon, B. f1avescens, in <strong>Canada</strong>; and<br />

two ichneumons, E. cintipes and Microplitis mediator, and one tachinid, E. consobrina,<br />

in western Europe. In <strong>Canada</strong>. one tachinid, A. cinerea, attacks the late instar larvae <strong>of</strong><br />

Mamestra cotlfigurata but no species <strong>of</strong> comparable effectiveness is reported from Eurasia.<br />

No egg parasitoids occur in <strong>Canada</strong> whereas T. evanescens occurs at generally low levels<br />

in Eurasia.<br />

The differences in the major components <strong>of</strong> the parasitoid guilds <strong>of</strong> M. configurata and<br />

<strong>of</strong> M. brassicae lead to the following recommendations for parasitoid introductions into<br />

<strong>Canada</strong>:<br />

(1) Parasitoids attacking the early larval ins tars<br />

(a) Ichneumons: B. f1avescetls in <strong>Canada</strong> is more effective than either E. cinctipes<br />

or Microplitis mediator in Eurasia. M. mediator does not go into diapause and does not<br />

occur in those parts <strong>of</strong> the Soviet Union where Mamestra brassicae is univoltine. Neither<br />

E. cinctipes nor Microplitis mediator should be considered for introduction into <strong>Canada</strong>.<br />

(b) Tachinids: Mamestra configurata does not have an effective species attacking the<br />

early larval instars as Mericia ampe/us is poorly adapted. E. consobrina is an important<br />

parasitoid in Eurasia. It successfully parasitizes Mamestra configurata in the laboratory<br />

(Wylie 1977a) and has potential as a biological control agent. Material for releases in<br />

<strong>Canada</strong> should be obtained from areas where M. brassicae is univoltine.<br />

(2) Parasitoids attacking late larval instars<br />

In <strong>Canada</strong>, A. ci1lerea is common, particularly during outbreaks, but no equivalent<br />

species is reported as a major parasitoid <strong>of</strong> M. brassicae in Eurasia. A. cinerea could<br />

be considered for introduction into the Soviet Union, perhaps as part <strong>of</strong> an exchange <strong>of</strong><br />

material involving E. C01lsobri1la. M. brassicae was a suitable host for A. cinerea in<br />

laboratory tests (Wylie 1977a).<br />

(3) Egg parasitoids<br />

The very low level <strong>of</strong> egg parasitism in M. brassicae does not suggest that T. evanescens<br />

would contribute significantly to the control <strong>of</strong> M. configurata in <strong>Canada</strong>. However, the<br />

local form <strong>of</strong> T. evanescens reported from the Irkutsk region (Kostylyova 1979) should<br />

be further investigated. If <strong>Canada</strong> should develop facilities for producing Trichogramma<br />

for inundative release, importation <strong>of</strong> the "Irkutsk" form would be desirable.<br />

Copies <strong>of</strong> the cited literature from the USSR were made available through the efforts<br />

<strong>of</strong> Mr A.1. Sylchenko, Ministry <strong>of</strong> Agriculture, Moscow, and Dr G. Tsubulskaja,<br />

Ukrainian Plant Protection Research Institute, Kiev. Their cooperation is greatly<br />

appreciated. Dr A.P. Arthur, Agriculture <strong>Canada</strong>, Saskatoon, kindly permitted the use<br />

<strong>of</strong> his data on parasitism <strong>of</strong> M. configurata in Saskatchewan.


Literature Cited<br />

Mamestra con/igurata Walker, 55<br />

Agarkov. V.M. (1974) Effect <strong>of</strong> entomophages on the abundance <strong>of</strong> the cabbage moth during'l period <strong>of</strong> low abundance. Bulletin <strong>of</strong>tht All·<br />

Union Scientific Re.rearch in.rtitllle for the PrOlectioll <strong>of</strong> Plants 27. 3-6 (Translated from Russian).<br />

Beirne. B.P. (1971) Pest insects <strong>of</strong> annual crop plants in <strong>Canada</strong>. Memoirs <strong>of</strong> the Enwmological <strong>Society</strong> <strong>of</strong> <strong>Canada</strong> 78. 124 pp.<br />

Central Intelligence Agency (1974) USSR agriculturallllias. U.S. Government Printing Office. 59 pp.<br />

Drozda. V.F.; Garnaga, N.G. (1978) Some biological and ecologic;11 char;scteristics <strong>of</strong> the t'lchinid Oy. Ernestia consobrina (Mg.). a parasite<br />

<strong>of</strong> the cabbage moth caterpill;sr. Scientific Pap'" <strong>of</strong> the Ukrainian Agrietlltural Academy 2(19. 23-25<br />

(Translated from Russian).<br />

Dzhurnn. V.N. (1979) The development <strong>of</strong> biological mcans for controlling the cabbage moth. Scientific Pap'" <strong>of</strong> the Ukraillian Agrimltural<br />

Academy 230. 25-26 (Translated from Russian).<br />

Hori. M. (1935) The cabbage moth (Barathra bra.uicae L.) in southern Saghalien. Report <strong>of</strong> the Saghalietr Central Experiment Station<br />

Sen-ice 1(3) (Translatcd from Japanese).<br />

Kop\;lIem. Kh.G. (1960) Parasites <strong>of</strong>the cabbage moth (Barathra brassicae L.) and the diamond· back moth (Pllllelia maculipennis Cun.) in<br />

the Moscow region. Entomologicl'l.'skol' Oboul'nil.' 39. 584-592.<br />

Kop\;lIem. Kh.G. (1962) Parasites <strong>of</strong> the cabbage moth and diamondback moth in the Moscow Region. Biological Methods <strong>of</strong> Controlling<br />

Pests and Discascs <strong>of</strong> Agricultural Crops I. 89-113 (Translated from Russian).<br />

Kostylyova. Ye.V. (1979) Experimental application <strong>of</strong> Trichogramma 'Igainstthe caggage moth in the Bratsk district <strong>of</strong> the Irkutsk region.<br />

In: Problems <strong>of</strong> Plant Protection in Eastern Siberia. Irkutsk. USSR. p. 34-39 (Translated from Russian).<br />

Prok<strong>of</strong>eva. N.A. (1976) Trichogramma in cabbage fields. Zaschita Rastenii 9. 19-20 (Translated from Russian).<br />

Razumov. V.P. (1972) Parasites <strong>of</strong> tbe cabbage moth (Maml'stra brassicae (L.)) in Gorky Oblasl. Proceedings <strong>of</strong> the Gorky Agricultural<br />

Institute 53.9-12 (Translated from Russian).<br />

Shchetinin. Yu.V. (1974) Parasites <strong>of</strong> the cabbage moth and diamond· back moth in the Tomsk region. In: The problems <strong>of</strong> entomology in<br />

Siberia. Novsibirsk. USSR. pp. 131·132 (Translated from Russian).<br />

SlabospilSkii. A.I. (1980) (Insect enemies <strong>of</strong> cabbage pests.) Zashchita Rastenii 5. 23 (in Russian).<br />

Stenin. V.I. (1974) The parasites. predators and polyhedrosis <strong>of</strong> the cabbage moth in the Kuban. Proceedings <strong>of</strong> the Kuibyshtl- Agricultural<br />

Institllle 125. 53-55 (Translated from Russian).<br />

Tsybul'ko. V.l. (1973) The main entomophages <strong>of</strong> the cabbage moth in Kharkov Province. Proceedings <strong>of</strong> the Kharkov Agricultural Institute<br />

182. 15-25 (Translatcd from Russian).<br />

Tumock. W.J.: Philip. H.G. (1979) The outbreak <strong>of</strong> benha armyworm. Maml'stra configurata (Noctuidae: Lepidoptera) in Albena 1971 to<br />

1975. Manitoba Entomologi.fl I I( 1977). 10-21.<br />

Wylie. H.G. (l977a) Observations on Athrycia cinerea (Diptera: Tachinidae). a parasite <strong>of</strong> Mamestra configurata (Lepidoptera: Noctuidae).<br />

Canadian Entomologist 109. 747-754.<br />

W)·lie. H.G. (1977b) Obsen'ations on Mericia ampe/us (Diptera: Tachinidae). an occasional parasite <strong>of</strong> hcnha armyworm. Mamestra<br />

configurata (Lepidoptera: Noctuidae) in western <strong>Canada</strong>. Canadian Entomologist 109. 1023-1024.<br />

Wylie. H.G. (1979) Insect parasites reared from artificial field populations ncar Winnipeg. Manitoba. Manitoba Entomologist 11(1977).<br />

SO-55.<br />

Wylie. H.G.; Bucher. G.E. (1977) The benha armyworm. Mamestra configurata (Lepidoptera: Noctuidae). Monalit), <strong>of</strong> immature stages on<br />

the rape crop. 1972-1975. Cantldian Entomologist 1119.823-837.


Blank Page<br />

56


Pest Status<br />

Background<br />

Chapter 16<br />

Manduca quinquemaculata (Haworth),<br />

Tomato Hornworm (Lepidoptera:<br />

Sphingidae)<br />

H.H. CHENG<br />

The tomato hornworm. Manduca quillquemaculata (Haworth). is principally a pest <strong>of</strong><br />

tomato and tobacco plants in <strong>Canada</strong>. The distribution <strong>of</strong> this species is North and South<br />

American. On tobacco. the first-instar and second-instar larvae make many small holes<br />

in the leaves, later they eat larger areas <strong>of</strong> the lamina. and by the last instar the larvae<br />

may strip the top and mid-leaves <strong>of</strong> the plants leaving only the midribs and main leaf<br />

veins. In <strong>Canada</strong>. a moderate to high population level <strong>of</strong> this pest may cause 0.5% to<br />

2.7% loss <strong>of</strong> the marketable tobacco in untreated fields (Cheng 1977a).<br />

The tomato hornworm has one complete generation and part <strong>of</strong> a second generation<br />

each year under southwestern Ontario climatic conditions. Only the first generation <strong>of</strong><br />

this species causes appreciable damage to tobacco. The insect overwinters as a pupa in<br />

the soil. Moth emergence occurs from late June to September. The adults fly at dusk and<br />

lay their eggs on the under side <strong>of</strong> the tobacco leaves during the same period. Eggs hatch<br />

within four to six days and the larvae dig into the soil to a depth <strong>of</strong> 10 to 12 cm and pupate.<br />

Most <strong>of</strong> the pupae remain in the soil until the following year. but some moths may<br />

emerge from late August to September to start a second generation. Ten-year blacklight<br />

trap studies showed that population levels <strong>of</strong> M. quillquemacu[ata in southwestern<br />

Ontario are generally low to moderate (unpublished data).<br />

The tomato hornworm can be successfully controlled using a number <strong>of</strong> different<br />

chemical insecticides, however. the use <strong>of</strong> chemical insecticides on tobacco presents a<br />

number <strong>of</strong> problems that may assume greater importance in the future. In particular.<br />

applications <strong>of</strong> certain insecticides on tobacco during late July and August result in<br />

insecticide residues (Cheng & Braun 1977). which have undesirable effects on the<br />

quality <strong>of</strong> tobacco and may put the Canadian tobacco export trade in serious jeopardy.<br />

Consequently. a search for and development <strong>of</strong> alternative control methods for hornworms<br />

has been carried out during the past 20 years. Among the more promising <strong>of</strong> the<br />

biological agents tested for the control <strong>of</strong> this pest are strains <strong>of</strong> the spore forming bacterium<br />

Bacillus thurillgiensLr Berliner. which was found to be as effective as chemical<br />

insecticides. and has been used successfully for controlling hornworms on tobacco<br />

(Guthrie et al. 1959. Creighton et al. 1961. Begg 1964. Bucher & Cheng 1971, Cheng<br />

1973. 1977b. 1978).<br />

Two commercial preparations <strong>of</strong> B. tllllrillgimsis, Dipe1® and Thuricide®-HPC, have<br />

been registered and recommended for use in controlling hornworms on tobacco since<br />

1973 (Anonymous 1973). A survey conducted in southwestern Ontario in 1977, 1978. and<br />

1979 indicated that about 20% <strong>of</strong> growers applied B. thuringiensis preparations for<br />

control <strong>of</strong> horn worms. When aphids and hornwormsoccurred in the same period, most <strong>of</strong><br />

the growers would apply chemical insecticides or a tank mix combination <strong>of</strong> B. thurillgiensis<br />

preparation for hornworm and a chemical insecticide for aphids because B. thuringiensis<br />

preparations 'gave almost no control <strong>of</strong> aphids.<br />

57


58 H. H. Cheng<br />

Evaluation <strong>of</strong> Control Attempts<br />

Recommendations<br />

Literature Cited<br />

The most important insect parasitoids <strong>of</strong> the tomato homworm are Apanteles spp.,<br />

which parasitize 10-20% <strong>of</strong> the hom worm larvae in tobacco fields each year.<br />

Apanteles spp. kill the fourth-instar and fifth-instar larvae and essentially prevent<br />

the homworm from causing maximum damage to the tobacco crop.<br />

Preparations <strong>of</strong> B. thllringiensis control tomato hornworms as well as chemical<br />

insecticides (Bucher & Cheng 1971. Cheng 1973. 1977b, 1978) and for approximately<br />

the same cost <strong>of</strong> $20 to $25 per hectare for the material. B. thllringiensis has the merits<br />

<strong>of</strong> being nontoxic to man. plants, and non-target animals including beneficial insects<br />

such as parasitoids and predators. Thus it should be used by growers when an<br />

economic population level <strong>of</strong> the tomato horn worm larvae occurs in the tobacco fields.<br />

Under the present system <strong>of</strong> annual cropping <strong>of</strong> Due-cured tobacco in Ontario,<br />

sequential applications <strong>of</strong> insecticides for homworm control and <strong>of</strong> sucker control<br />

agents within short periods are common practices in tobacco production. To reduce the<br />

cost <strong>of</strong> control measures, attempts have been made to mix recommended insecticides<br />

and sucker control agents in the same tank and apply them to tobacco plants for<br />

control <strong>of</strong> hornworms and suckers in one spray operation. Five year tests indicated that<br />

all recommended insecticides used for homworm control can be mixed with sucker<br />

control agents in the same tank before application without losing their efficacy and<br />

caused no effects on the yield and quality <strong>of</strong> Due-cured tobacco in Ontario (Cheng 1977b,<br />

1978, 1980).<br />

At the Delhi Research Station, monitoring and economic threshold data have been<br />

developed for M. qllinquemaculata (Cheng 1977a). Since this pest overwinters as a<br />

pupa in the soil, early detection methods involve the capture <strong>of</strong> moths in light traps.<br />

Larval densities are monitored from mid-July to mid-August and insecticide applications<br />

are recommended when 5 or more large larvae (larger than 3 cm) are found per 100<br />

plants. Generally. M. qllinqllemacillata can be controlled for the whole growing season<br />

by a single application <strong>of</strong> B. thllringiensis or one <strong>of</strong> the recommended chemical<br />

insecticides.<br />

Detailed studies on the bionomics <strong>of</strong> the tomato hornworm in Ontario should be<br />

conducted. Although preparations <strong>of</strong> B. thuringiensis have provided excellent control <strong>of</strong><br />

the hornworm larvae, the rapid drop <strong>of</strong> larvae from plants treated with chemical insecticides<br />

is much more dramatic than the cessation <strong>of</strong> feeding and the slow death <strong>of</strong> B.t.infected<br />

larvae. Growers, therefore, should be better informed <strong>of</strong>the merits <strong>of</strong> B. thuringiensis<br />

to <strong>of</strong>fset their natural preference for a chemical insecticide that has more<br />

spectacular effects. A similar approach should be made in other areas <strong>of</strong> <strong>Canada</strong> where<br />

the tomato hornworm is a problem.<br />

Anonymous (1973) 1973 Tobacco production recommendations. Ontario Ministry Agriculture and Food, Pub!. 298. 32pp.<br />

Begg, J.A. (1964) Microbial and chemical control <strong>of</strong> homworms attacking tobacco in Ontario. Journal <strong>of</strong> Economic Entomology 57, 646-649.<br />

Bucher, G .E.; Cheng, H. H. (1971) Comparison <strong>of</strong> Bacillus Ihuringiensis preparations with carbaryl for homworm (Lepidoptera: Sphingidae)<br />

control on tobacco. Canadian Entomologist 103, 142-144.<br />

Cheng. H.H. (1973) Microplot test using microbial and chemical insecticides for control <strong>of</strong> tomato homworms on tobacco in Ontario. The<br />

Lighter 43(3). 10-13.


Mtmduca qllillqllemaclliata (Haworth), 59<br />

Cheng, H.H. (1977a) F1ue-cured tobacco losses caused by the tomato hornworm, Manduca quinquemaculala (Lepidoptera: Sphingidae), at<br />

various infestation levels in Ontario. Canadian Entomologist 109,1091-1095.<br />

Cheng, H.H. (1977b) Insecticides and sucker control chemicals: compatibility and effects on green peach aphid, tomato hornworm and<br />

suckers on flue-cured tobacco in Ontario. Tobacco Science 21, 108-111.<br />

Cheng, H.H. (1978) Pesticides play important role. Canadian Tobacco Grower 26(8),31-33.<br />

Cheng, H.H. (1980) Erfects <strong>of</strong> tank-mixed combinations <strong>of</strong> insecticides and sucker control agents on efficacy and on the yield and quality <strong>of</strong><br />

flue-cured tobacco. Proceedings <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> Ontario Ill, 7 -12.<br />

Cheng, H.H.; Braun, H.E. (1977) Chlorpyrifos, carbaryl, endosulfan, leptophos, and trichlorfon residues on cured tobacco leaves from fieldtreated<br />

tobacco in Ontario. Canadian 10urnal <strong>of</strong> Plant Science 57,689-695.<br />

Creighton, C.S.; Kinard, W.S.; Allen, N. (1961) Effectiveness <strong>of</strong> Bacillus tlluringiensis and several chemical insecticides for control <strong>of</strong><br />

budworms and hornworms on tobacco. 10urnal <strong>of</strong> Economic Entomology 54, 1112-1114.<br />

Guthrie, F.E.; Rabb, R.L.; Bowery, T.G. (1959) Evaluation <strong>of</strong> candidate insecticides and insect pathogens for tobacco hornworm control,<br />

1956-1958. 10urnal <strong>of</strong> Economic Entomology 52, 798-804.


Blank Page<br />

60


Chapter 17<br />

Melanoplus spp., Camnula pellucida<br />

(Scudder), and other Grasshoppers<br />

(Orthoptera: Acrididae)<br />

A.B. EWEN and M.K. MUKERJI<br />

Grasshoppers have been important destructive pests since crop production began on the<br />

Canadian prairies. Four species: the migratory grasshopper. Melanoplus sanguillipes<br />

(Fab.); the twostriped grasshopper, M. bivittatllS (Say); the Packard grasshopper, M.<br />

packardii Scudder; and the c1earwinged grasshopper, Camnllia pellucida (Scudder);<br />

generally arc considered to be the most important. Natural enemies and weather<br />

conditions cause marked fluctuations in their populations from year to year from one<br />

area to another, and several population peaks have been recorded during the more than<br />

thirty years that grasshopper surveys have been conducted in Saskatchewan (Gage &<br />

Mukerji 1977). High populations and significant damage to crops have been associated<br />

with early egg hatch due to an early, warm, and dry spring (Randell & Mukerji 1974).<br />

Grasshoppers host a variety <strong>of</strong> pathogenic micro-organisms including fungi, bacteria,<br />

viruses, and protozoans. Epizootics <strong>of</strong> fungal pathogens can produce drastic reductions<br />

in grasshopper populations. Pickford & Riegert (1964) showed that Emomophlhora<br />

grylli Fres. was a major factor responsible for significant reductions in grasshopper<br />

populations, espciaJly C. pellucida, in Saskatchewan in 1963. Although there have<br />

been other incidents <strong>of</strong> fungal epizootics, these pathogens have not given consistent<br />

control <strong>of</strong> grasshopper populations because <strong>of</strong> their restrictive requirements <strong>of</strong> specific<br />

moisture and temperature conditions for initiation and dissemination <strong>of</strong> infection in the<br />

host populations (Henry 1970). Most bacteria associated with grasshoppers are<br />

facultative pathogens which cannot multiply in the host gut but require direct passage<br />

into the haemocoele (Bucher 1960). Also. bacteria and many <strong>of</strong> the fungi pathogenic in<br />

grasshoppers are not suitable microbial control agents because they are associated<br />

with or pathogenic in other insects and animals (Henry 1970).<br />

Viral and protozoan pathogens have received more attention as potential biological<br />

control agents for grasshoppers. An entomopox virus (Henry & Jutila 1966) and a<br />

crystalline-array virus <strong>of</strong> the picornovirus group (Jutila et al. 1970) have been isolated<br />

from M. sangllinipes and M. bivittatus, respectively. Although these viruses arc<br />

uncommon in grasshoppers under normal conditions, observations indicate that they<br />

may be capable <strong>of</strong> lowering the population density <strong>of</strong> their hosts. Prolonged effects <strong>of</strong><br />

these pathogens have yet to be assessed. Of the protozoans, Malameba locustae (King<br />

& Taylor), several species <strong>of</strong> grcgarincs, three species <strong>of</strong> Nosema (N. locuslae Canning.<br />

N. acridophagus Henry, and N. cuneatum Henry), and a number <strong>of</strong> other microsporidians<br />

have been reported or described in grasshoppers. Malameba locustae and the<br />

gregarines (the most common protozoans in grasshoppers) cause chronic infections but<br />

have little immediate effect on their hosts.<br />

Most research effort in recent years has been directed towards using the Nosema<br />

spp. as biological control agents for the management <strong>of</strong> grasshopper populations and,<br />

<strong>of</strong> these. N. loclIstae has received the most attention. In 1980 a quantitative study <strong>of</strong><br />

the effect <strong>of</strong> N. /ocuslae as a control agent against grasshopper populations was<br />

carried out in a short grass pasture in east central Saskatchewan (Ewen & Mukerji<br />

1980). The predominant grasshopper species were M. sangllinipes, M. packardii. and<br />

C. pellucida, and about 50% <strong>of</strong> these populations were infected between 4 and 5 weeks<br />

61


62 A. B. Ewen and M. K. Mukcrji<br />

Literature Cited<br />

after application <strong>of</strong> 2.5-5.0 .x 10' spores <strong>of</strong> N. iocustae per ha. Maxima <strong>of</strong> 95-100%<br />

infection were evident between 9 and 12 weeks after application for all three<br />

grasshopper species. This increase in prevalence <strong>of</strong> infection throughout the season<br />

probably was due to the fact that feeding grasshoppers tend to consume grasshopper<br />

cadavers that they find, and data (Ewen & Mukerji 1980) indicate that most cadavers<br />

would contain mature spores <strong>of</strong> N. iocllStae later in the season.<br />

M. sanguinipes showed the greatest vulnerability to N. iocllStae infection among the<br />

three predominant grasshopper species in the Saskatchewan study (Ewen & Mukerji<br />

1980). At a 50% level <strong>of</strong> infection. its population was reduced 26%, significantly more<br />

than either M. packardii (20%), or C. pellucida (8%). Overall, the treated populations <strong>of</strong><br />

M. sanguinipes and M. packardii were reduced by about 20%, 4 weeks after spore<br />

application and by about 60%,8 weeks later. On the other hand, the population <strong>of</strong> C.<br />

pellucida, although showing about the same prevalence <strong>of</strong> infection as the MeianopillS<br />

populations, was reduced overall by only 28%. 12 weeks after application <strong>of</strong> N. iocllStae<br />

spores. It seems that the Cyrtacanthacridinae are more vulnerable to N. iocllStae<br />

infection than are the Oedipodinae.<br />

Grasshoppers that survived to lay eggs appeared to be infected to the extent that their<br />

fecundity was reduced. This was especially true <strong>of</strong> M. packardii, where the females laid<br />

about 65% fewer eggs than untreated populations (Ewen & Mukerji 1980). Egg production<br />

by the treated M. sanguinipes populations also was reduced markedly when<br />

compared to the control populations. Not enough data were collected to evaluate the<br />

effect <strong>of</strong> the pathogen on fecundity <strong>of</strong> C. pellucida. Data show that mature spores <strong>of</strong> N.<br />

iocllstae can overwinter (presumably in cadavers,) but it is not known if sufficient<br />

numbers survive to adequately infect the next year's populations.<br />

Applications <strong>of</strong> N. iocllStae spores in Saskatchewan were made to grasshopper<br />

populations that were predominantly in the third nymphal instar (Ewen & Mukerji 1980).<br />

Applications to younger populations resulted in the insects being killed at much lower<br />

levels <strong>of</strong> infection. and this could have important implications for grasshopper control in<br />

crop land. Crop defoliation could be reduced by killing the grasshoppers when they were<br />

first-instar or second-instar nymphs. but then fewer spores would be produced for infection<br />

<strong>of</strong> other grasshoppers. Thus. reducing the grasshopper populations earlier would<br />

better protect the crop. but a second application <strong>of</strong> spores a week or two later would be<br />

necessary because considerable recruitment would still be possible in these younger<br />

populations.<br />

Bucher. G.E. (1960) Potcntial bacterial pathogens <strong>of</strong> insects and their characteristics. Journal <strong>of</strong> Insect Pathology 2. 172-195.<br />

Ewen. AI B.; Mukerji. M.K. (1980) Evaluation <strong>of</strong> Nmema locus/ae (Microsporidia) as a control agent <strong>of</strong> grasshopper populations in<br />

Saskatchewan. JOIlrtral <strong>of</strong> Invertebrate Pathology 35. 295-303.<br />

Gage. S.H.; Mukerji. M.K. (1977) A perspective <strong>of</strong> grasshopper distribution in Saskatchewan and interrelationship with weather.<br />

Environmental Entomology 6. 469-479.<br />

Henry. J.E. (1970) Potential microbial control <strong>of</strong> grasshoppers and locusts. Proceedings <strong>of</strong> the International Study Conference. Current &<br />

Future Problems <strong>of</strong> Acridology. pp. 465-468.<br />

Henry. J.E.; Jutila. J.W. (1966) The isolation <strong>of</strong> a polyhedrosis virus from a grasshopper. Journal <strong>of</strong> Invertebrate Pathology 8.417-418.<br />

Jutila. J.W.; Henry. J.E.; Anacker. R.L.; Brown. W.R. (1970) Some properties <strong>of</strong> a crystallinc-array virus (CAV) isolated from the<br />

grasshopper Melanoplus biviltatus (Say) (Orthoptera: Acrididae). Journal <strong>of</strong> invertebrate Pathology 15.<br />

225-23\.<br />

Pickford. R.; Riegert. P. W. (1964) The fungous disease eaused by Entomophthora grylfi Fres .• and its effects on grasshopper populations in<br />

Saskatchcwan in 1963. Canadian Enlomologist 96. 1158-1166.<br />

Randen. R.L.; Mukcrji. M. K. (1974) A technique for estimating hatching <strong>of</strong> natural egg populations <strong>of</strong> Melanoplus sanguinipes (Orthoptera:<br />

Acrididae). Canadian EnlOmologift 106. Sot -SI2.


Chapter 18<br />

Musca domestica L., House Fly (Diptera:<br />

Muscidae)<br />

R.A. COSTELLO<br />

Aies, particularly Musea domes/iea L., associated with egg producing operations have<br />

become an increasing problem in British Columbia's Fraser Valley in recent years.<br />

Conflicts between farmers and their nonfarming neighbours have intensified with the<br />

encroachment <strong>of</strong> urban development on agricultural areas. Flies can also be a considerable<br />

annoyance to farm operators and their families and employees.<br />

Certain parasitoid wasps have been found effective in suppressing nuisance fly populations<br />

in egg producing operations in California (Legner & Dietrick 1972, Legner e/ al.<br />

1975). A commercial insectary in that state provided the parasitoids for this study. The<br />

pupal parasitoid Spa/angia endius Walker (Hymenoptera: Pteromalidae) was selected<br />

for this trial because <strong>of</strong> its tendency to burrow deeper into manure than other available<br />

parasitoids in search <strong>of</strong> pupae (Legner 1977). Pre-release sampling indicated that more<br />

than 90% <strong>of</strong> the pupae were further than 3 cm below the surface.<br />

In 1978, a total <strong>of</strong> 400000 Spa/angia endius, in the form <strong>of</strong> parasitized house fly pupae,<br />

were released into a deep pit layer barn in Aldergrove, British Columbia. The barn,<br />

measuring 91 m by 12 m, contained 27 000 layers. The pit contained a 9 month accumulation<br />

<strong>of</strong> manure at the start <strong>of</strong> the trial. Four releases <strong>of</strong> 100 000 parasitoids each were<br />

made at 3 week intervals commencing 20 April. No other control methods were applied<br />

in this barn. Fly numbers, as indicated by "spot cards" (Rutz & Axtell 1979). were<br />

compared to those in a bam 30 m away where fly control consisted <strong>of</strong> chemical insecticides<br />

applied as space sprays. residual surface sprays, and floor baits. The spot cards<br />

consisted <strong>of</strong> white 7.6 x 12.7 cm unlined file cards. These were pinned to posts in the<br />

barns, collected weekly. and the fecal and regurgitation marks counted.<br />

Levels <strong>of</strong> parasitism were measured by collecting manure samples once a week.<br />

floating out fly pupae, and holding them at 30"C until either flies or parasitoids emerged.<br />

This also provided an indication <strong>of</strong> the fly species involved and all pupae collected were<br />

Musca domes/iea.<br />

Ay numbers were monitored in both barns from 25 April to 13 June. The spot card<br />

counts were similar for the first weeks, averaging 45.3 per card in the bam using<br />

chemical control methods and 35.7 in the barn where parasitoids were released. By the<br />

end <strong>of</strong> the third week, mean spot card counts rose to 390.2 in the chemical control barn<br />

and fell to 26.9 in the barn containing Spa/angia endius. Repeated applications <strong>of</strong><br />

pyrethrum as a thermal fog resulted in a reduction <strong>of</strong> the mean spot card count to 213.2<br />

by the end <strong>of</strong> the 6th week when the manure was removed and spot card counts were<br />

discontinued. Spot counts in the barn using parasitoids averaged 20 to 25 during the 4th,<br />

5th and 6th weeks <strong>of</strong> the trial.<br />

Six weeks after the initial introduction <strong>of</strong> parasitoids 63% <strong>of</strong> the fly pupae separated<br />

from manure samples contained Spa/angia endius larvae and pupae. This was the highest<br />

level <strong>of</strong> parasitism found during the sampling period.<br />

House fly numbers in the barn containing parasitoids rcmnincd nt acceptnble levels<br />

throughout the trial and the remainder <strong>of</strong> the fly senson, while flies in the bam employing<br />

chemical control methods were intolerably abundant and eventually necessitated<br />

removing the manure to relieve the problem. The commercial value <strong>of</strong> the parasitoids<br />

used in this study was approximately $800, about the same as the operator spent on<br />

chemicals the previous year.<br />

63


64 R. A. Costello<br />

Literature Cited<br />

Further laboratory and field studies are being carried out to develop an integrated<br />

approach to fly control in egg layer barns entailing parasitoid introductions, manure<br />

management, and chemical insecticides.<br />

Legner. E.F. (1977) Temperature. humidity and depth <strong>of</strong> habitat influencing host destruction and fecundity <strong>of</strong> muscoid fly parasites.<br />

Entomophaga 22. 199-206.<br />

Legner. E.F.; Dietrick. E.J. (1972) Inundation with parasitic insects to control filth breeding flics in California. Proceed;II8s o/the CalifornitJ<br />

Mosquito Control Association 40, 129-130.<br />

Legner. E.F.; Bowen. W.R.; Rooney. W.F.; McKeen, W.O.; Johnston. G.W. (1975) Integrated fly control on poultry ranches. California<br />

Agriculture 29(5). 8-10.<br />

RUll. D.A.; Aldell, R.C. (1979) Sustained releases <strong>of</strong> Muscidifurru raptor (Hymenoptera: Pteromalidae) for house fly (Musca domestica)<br />

control in two types <strong>of</strong> caged·layer poultry houses. Environmellta/ Entomology 8. 1105-1110.


Pest Status<br />

Background<br />

Chapter 19<br />

Releases and Recoveries in Ontario<br />

Oulema melanopus (L.), Cereal<br />

Leaf Beetle (Coleoptera: Chrysomelidae)<br />

D.G. HARCOURT, J.e. GUPPY and C.R. ELLIS<br />

During the early and mid 1970s, the cereal leaf beetle (eLB), Oulema melanopus (L.),<br />

began to threaten the production <strong>of</strong> small grains in Ontario. Following its initial<br />

discovery in Essex County near the border <strong>of</strong> Michigan in 1965, it spread rapidly<br />

eastward and by 1975 it had occupied most <strong>of</strong> Ontario south <strong>of</strong> Hwy 17. Populations<br />

reached economic levels in 1973, and in 1974 chemical treatment was required in ca. 75<br />

fields within the triangular area bounded by the counties <strong>of</strong> Lincoln, Grey, and<br />

Durham (Bereza 1974). In 1976, it occurred in damaging numbers on Manitoulin Island<br />

and, in 1977, it was found north <strong>of</strong> Lake Huron in the districts <strong>of</strong> Algoma, Sudbury,<br />

and Nipissing (Ellis et al. 1979).<br />

Like many <strong>of</strong> our introduced pests, the cereal leaf beetle is rarely troublesome in its<br />

native Europe where it is attacked by a complex <strong>of</strong> natural enemies. One <strong>of</strong> these, the<br />

larval cndoparasitoid Tetrastichus julis (Walker) was introduced against the beetle by<br />

research workers in Michigan in 1967. It was recovered in low numbers in 1969 (Stehr<br />

1970) and was extensively recolonized in 1971, when further releases were made in<br />

Michigan, Indiana, Ohio, West Virginia, and New York (Dysart et 01. 1973).<br />

T. julis is mostly bivoltine in Michigan (Gage & Haynes 1975). It emerges from its<br />

overwintering sites in late May and flies to nearby grain fields where it deposits its eggs<br />

in the first host larvae that appear. The gregarious parasitoid larvae grow slowly within<br />

their host while it is feeding. As soon as the host larva enters the soil and forms its<br />

pupal cell, the parasitoids quickly complete their development and form naked pupae<br />

within the cell. Diapause is facultative, but the incidence increases as the season<br />

progresses. Some adults emerge in early summer and give rise to a second generation.<br />

The remainder <strong>of</strong> the first generation and all <strong>of</strong> the second overwinter as last instar<br />

larvae inside the pupal cells <strong>of</strong> their hosts. In Michigan, roughly 80% <strong>of</strong> the population<br />

enters a second generation; this generation is poorly synchronized with populations <strong>of</strong><br />

its host since at the time <strong>of</strong> emergence relatively few beetle larvae are available for the<br />

adults to attack (Gage & Haynes 1975).<br />

Studies in Michigan during the early seventies indicated that T. julis had poor<br />

powers <strong>of</strong> dispersal (Haynes et -al. 1974). Natural spread was slow and rates <strong>of</strong><br />

parasitism in the United States midwest remained well below that needed to maintain<br />

populations at subeconomic levels.<br />

Large-scale releases <strong>of</strong> T. julis were made at 4 sites in southcentral Ontario in June<br />

1974 when parasitized host larvae were set out in fields <strong>of</strong> oats and barley (Harcourt et<br />

al. 1977). This has been the only parasitoid released in the province.<br />

In 1975, life table studies <strong>of</strong> the host in the Quinte area <strong>of</strong> southern Ontario showed<br />

that 55% and 43% respectively <strong>of</strong> the pupal cells in study plots in Northumberland and<br />

65


66 D. G. Harcourt, J. C. Guppy and C. R. Ellis<br />

Table 10<br />

Current Status <strong>of</strong> T. julis<br />

Hastings Counties (Harcourt et al. 1977) contained larvae <strong>of</strong> T. julis. These findings,<br />

which were unexpected in light <strong>of</strong> the Michigan experience, prompted closer examination<br />

<strong>of</strong> larval collections made during the course <strong>of</strong> annual CLB population surveys by Ontario<br />

Ministry <strong>of</strong> Agriculture and Food personnel at 8 sites within the province during 1975.<br />

The surveys consisted <strong>of</strong> 250 net sweeps from each <strong>of</strong> 30 fields <strong>of</strong> spring grain located<br />

within a 75 km! area. Eight such areas were sampled in late June following the<br />

accumulation <strong>of</strong> 475°D>9°C.<br />

Table 10 shows that T. julis had dispersed widely across Ontario by 1975. Rates <strong>of</strong><br />

parasitism were highest in southwestern Ontario, lowest in eastern Ontario, and<br />

intermediate in other parts <strong>of</strong> the province. Overall parasitism was 84%. The number<br />

<strong>of</strong> parasitoids per host ranged from 1 to 44 and averaged 5.9. The four host instars were<br />

parasitized in roughly the same proportion (75, 87, 84,84), suggesting that the larvae<br />

were attacked when small.<br />

Parasitism <strong>of</strong> the cereal leaf beetle,<br />

(Walker) in Ontario, 1975.<br />

County Township No. host<br />

larvae collected<br />

Oulema melanopus (L.), by Tetrastiehus julis<br />

Kent Merlin 894 95.0<br />

Middlesex Glencoe 212 91.9<br />

Essex Malden 526 88.6<br />

Wellington Fergus 227 83.7<br />

York Whitchurch 143 83.2<br />

Ontario Brock 15 73.3<br />

Bruce Mildmay 156 68.6<br />

Renfrew Bromley 147 15.0<br />

TOTALS 2320 84.4<br />

% parasitized<br />

Sweepnet samples <strong>of</strong> spring grain on Manitoulin Island in early July 1976 showed that<br />

the parasitism rate was 87% (Ellis et al. 1979).<br />

Since 1974, there have been no reports <strong>of</strong> economic damage by the cereal leaf beetle in<br />

Ontario other than on Manitoulin Island. In most fields <strong>of</strong> spring grain it has been<br />

difficult to find larvae. This raised fears that the host population might become too low to<br />

maintain the parasitoid which would set the stage for a resurgence <strong>of</strong> the pest. For this<br />

reason, extensive sampling has been carried out in representative sites in Ontario since<br />

1977. Rates <strong>of</strong> parasitism in the past 4 years have ranged from 31-68%, indicating that T.<br />

julis has the capacity to maintain itself as an effective biological control agent at very low<br />

host densities. There has been no indication <strong>of</strong> resurgence <strong>of</strong> CLB populations and no<br />

damage from the pest was reported in Ontario during the 1977-80 period.<br />

Because the distribution <strong>of</strong> T. julis in 1975 was more extensive than the area <strong>of</strong> the 1974<br />

releases and because its population density at most sites was inexplicably high, it is<br />

concluded that the parasitoid spread into Ontario as a result <strong>of</strong> the earlier releases in<br />

Michigan and other parts <strong>of</strong> the United States. It is also clear that the parasitoid has<br />

better powers <strong>of</strong> dispersal than originally reported.


Recommendations<br />

Literature Cited<br />

Ou/ema me/at/opus (L.). 67<br />

We believe that there are two reasons for the spectacular success <strong>of</strong> T. iulis in Ontario<br />

relative to that in Michigan. First, the larval period <strong>of</strong> the cereal leaf beetle occurs after<br />

the seasonal accumulation <strong>of</strong> 235°0>9°C (Fulton & Haynes 1975) throughout its range,<br />

and this is therefore later in Ontario than in the United States midwest. Since the<br />

developmental threshold <strong>of</strong> T. iu/is resembles that <strong>of</strong> its host (Gage & Haynes 1975), it<br />

similarly emerges later in its northern range and its larvae complete their development<br />

later in the season. Hence, a greater proportion <strong>of</strong> the first generation enter diapause.<br />

Second, in deference to the harsher climate than in the United States midwest, agricultural<br />

practice in most parts <strong>of</strong> Ontario favours spring grains as opposed to winter wheat.<br />

These frequently grow in rotation with, and as a nurse crop for, alfalfa and other<br />

legumes. Hence, the field is left untilled after harvest and host pupal cells below soil<br />

level are not disturbed by plowing and disking, a practice which can destroy more than<br />

95% <strong>of</strong> the parasitoid population (Haynes et al. 1973).<br />

In view <strong>of</strong> the potential importance <strong>of</strong> the cereal leaf beetle to culture <strong>of</strong> spring grains in<br />

<strong>Canada</strong>, rates <strong>of</strong> attack <strong>of</strong> T. iulis relative to numbers <strong>of</strong> its host should be monitored at<br />

regular intervals throughout its range.<br />

Bereza. K. (1914) Ontario insects. Canadian Agricultural/nseet Pest Review 52. 16.<br />

Dysan. RJ.; Maltby. H.L.; Brunson. M.H. (1913) Larval parasites <strong>of</strong> Oulema melanopus in Europe and their colonization in the United<br />

States. Entomophaga 18. 133-161.<br />

Ellis. e.R.; Harcoun. D.G.; Dubois-Martin. D. (1919) The current status in Ontario <strong>of</strong> Tetrastichus julis (Hymenoptera: Eulophidae). a<br />

parasitoid <strong>of</strong> the cereal leaf beetle. Proceedings <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> Ontario 109. 23-26.<br />

Fulton. W.e.; Haynes. D.L. (1915) Computer mapping in pest management. Environmental Entomology 4.357-360.<br />

Gage. S. H.; Haynes. D. L. (1975) Emergence under natural and manipulated conditions <strong>of</strong> Tetrastichus julis. an introduced larval parasite <strong>of</strong><br />

the cereal leaf beetle. with rcCerence to regional population management. Environmental Entomology 4.<br />

425-434.<br />

Harcourt. D. G.; Guppy. J .C.; Ellis. C. R.E. (1917) Establishment and spread <strong>of</strong> Tetrastiehus julis (Hymenoptera: Eulophidae) a parasitoid <strong>of</strong><br />

the cereal leaf beetle in Ontario. Canadian Entomologist 109, 473-476.<br />

Haynes. D.L.; Brandenburg. R.K.; Fisher, P.D. (1913) Environmental monitoring network for pest management systems. Environmental<br />

Entomology 2. 889-899.<br />

Haynes. D.L.; Gage. S.H.; Fulton. W. (1974). Management <strong>of</strong> the cereal leaf beetle pest ecosystem. Quaestiones Entomologieae 10,<br />

165-176.<br />

Stehr. F. W. (1910) Establishment in the United States <strong>of</strong> Tetrastiehus julis. a larval parasite <strong>of</strong> the cereal leaf beetle. Journal <strong>of</strong> Economic<br />

Entomology 63.1968-1969.


Blank Page<br />

68


Pest Status and Background<br />

Chapter 20<br />

Phyllonorycter blancardella (F.), Spotted<br />

Tentiform Leafminer (Lepidoptera:<br />

Gracillariidae)<br />

J.E. LAING<br />

The spotted tentiform leafminer (STLM). Phyllonorycler blancardella (F.). is an indirect<br />

pest <strong>of</strong> apple in eastern <strong>Canada</strong> and the United States. This species probably<br />

entered North America from Europe early in the 1900s. Severe infestations (> 20 mines<br />

per leaf) may cause early leaf drop. stunting <strong>of</strong> terminal growth. premature fruit<br />

ripening, and fruit drop with a resultant decrease in yield and reduction <strong>of</strong> fruit set the<br />

year following the infestation (Pottinger & LeRoux 1971. Johnson 1975). There are three<br />

generations <strong>of</strong> the leafminer per year throughout its range in North America. The<br />

leafminer overwinters as a pupa. emerges in May. and begins to oviposit first-generation<br />

eggs. Adults <strong>of</strong> the first generation are present in orchards from mid-June through July.<br />

Second-generation adults appear in August (Johnson 1975. Laing 1977).<br />

Although many species <strong>of</strong> parasitoids have been recorded from STLM (Pottinger &<br />

Le Roux 1971. Johnson el al. 1978). this species is able to reach damaging levels in some<br />

orchards each year by the third or overwintering generation. Because <strong>of</strong> the cryptic<br />

feeding habits <strong>of</strong> the larvae. chemical controls must be either systemic or timed to<br />

coincide with adult emergence. The latter has been facilitated by the synthesis <strong>of</strong> a sex<br />

attractant for STLM (Roel<strong>of</strong>s el al. 1977) and the development <strong>of</strong> heat unit requirements<br />

above a threshold <strong>of</strong> approximately 7°C for emergence <strong>of</strong> the adults (Johnson el 01.<br />

1979). Although individual growers may obtain good control <strong>of</strong> STLM at times. reinvasion<br />

<strong>of</strong> commercial orchards likely occurs from hedgerow and abandoned or poorly<br />

maintained orchards.<br />

The following parasitoids have been reported from STLM in Ontario and Quebec<br />

(Pottinger & LeRoux 1971. Johnson el al. 1978):<br />

Braconidae Apanteles ornigis Weed<br />

Eulophidae Sympiesis marylandensis Girault<br />

Sympiesis bimaculatipennis Girault<br />

Sympiesis sericeicornis (Nees)<br />

Sympiesis conica (Provancher)<br />

Pnigalio maculipes (Crawford)<br />

Pnigalio flavipes (Ashmead)<br />

Pnigalio uroplatae (Howard)<br />

Chrysocharis cuspidogaster (Yosh.)<br />

Chrysocharis Iricinclus Ashm.<br />

Chrysocharis laricinellae (Ratz.)<br />

Tetrastichus sp.<br />

Cirrospilus cinclilhorax (Girault)<br />

Zagrammosoma mullilinealum (Ashmead)<br />

Horismenus Jraternus (Fitch)<br />

Eupelmidae Eupelmella vesicularis (Retzius)<br />

The most abundant parasitoids <strong>of</strong> STLM are the braconid Apanteles ornigis Weed in<br />

the first and third generations. and the chalcids Sympiesis marylandensis Gir. and S.<br />

69


Ph yJlollorycter blallcardelltr (F.) , 71<br />

Johnson. E.F .• Trouier. R.: L.'1ing. J.E. (1979) Degree-day relationships to the development <strong>of</strong> Lilhocollelis blancardella (Lepidoptem:<br />

Gracillariidae) and its parasite Apenlldes mnigis (Hymenoptera: Braconidae). Canadiall EtJlonrologisl<br />

III, II77- IIs...<br />

Laing. J.E. (l9n) Spotted tentifonn Icafminer. Ontario Ministry <strong>of</strong> Agriculture and Food, Agdex Factshcct. 3 pp.<br />

Laing. J.E.; Her:lty. J.M. (1981) Establishment in <strong>Canada</strong> <strong>of</strong> the parasite "'pallldes pedias Nixon on the spotted lenliform leafminer<br />

Phyllonorycler blancardella (Fabr.). EIII'irtmmelllal Emomology 10, 933-935.<br />

Nixon, G.E.J. (1973) A revision <strong>of</strong> Ihe north-western Europc.1D species <strong>of</strong> Ihe "ilripemlis, pallipes, octonariw, triallgulator, fratermlS,<br />

foniosllS, parasilellae, melacarpalis, and cirC/tt1lscripIILf-groups <strong>of</strong> ApatJIeles Forster (Hymenoptera:<br />

Braconidae)_ Bullelill <strong>of</strong> EIlIt}IJwltlgictz/ Research 63, 169-228.<br />

POllinger, R. P.; Le Roux, E.J. (1971) The biology and dynamics <strong>of</strong> Lil/mcolletis IJIclllcelrdelia (Lepidoptera: Gracillariidae) on apple in<br />

Quc!bec. Memoirs <strong>of</strong> Ihe ElllolllologiclIl Sociely <strong>of</strong> Calltlda 77, 437 pp.<br />

Roel<strong>of</strong>s, W.L.: Reissig, W.H. Weires, R.W. (1977) Sex auractalll for the spoiled tentiform learminer moth Lithocolletis b/allcardelltl.<br />

Environlllellla/ Etllolllolog}, 6, 373 - 374.<br />

SWilD, D.1. (1973) Evalualion <strong>of</strong> biologicnl conlrol <strong>of</strong> Ihe ouk Icafminer I'lzyllrmor},ctermessalliella (Zeit) (Lepidoptera: Gracillnriidae) in<br />

New Zealand. Bulletill <strong>of</strong> Elllr1lTwlogiCll/ Re.rI!rlrclz 63, 49-55.


Blank Page<br />

72


Pest Status<br />

Background<br />

Chapter 21<br />

Phyllotreta spp., Flea Beetles (Coleoptera:<br />

Chrysomelidae)<br />

H.G. WYLIE, W.l. TURNOCK and L. BURGESS<br />

Damage to rape crops in western <strong>Canada</strong> by flea beetles is caused primarily by adults <strong>of</strong><br />

two species, Phy/lotreta crllciferae (Goeze) and Phyl/otreta strio/ata (F.). A third flea<br />

beetle, Psylliodes puncru/ata Melsh., occurs in virtually all seedling rape crops on the<br />

prairies but by itself does not constitute a serious problem. Phyllolrela robusta Lee. and<br />

Phyllolrela a/bionica (Lee.) occur occasionally on rape crops in Saskatchewan (Burgess<br />

1977a) and three other described crucifer-feeding species in the genus Phyllotreta are<br />

known to occur in that province (Burgess 1981b).<br />

P. cruciferae was introduced on the west coast <strong>of</strong> North America in the early 1920s<br />

(Milliron 1953) and was abundant on crucifers in the Prairie Provinces by the late 1930s<br />

and early 1940s (Westdal & Romanow 1972). P. cruciferae has usually been the<br />

dominant flea beetle in rape fields across the prairies, particularly in the more southerly<br />

parts <strong>of</strong> the rape-growing area.<br />

P. slrio/ala apparently was introduced to North America before 1801 (Chittenden<br />

1923) and was present in <strong>Canada</strong> from the Atlantic provinces to British Columbia by the<br />

early 1900s (Burgess 1977a). This species infests rape crops across the parkland<br />

agricultural area from southern Manitoba to north <strong>of</strong> the Peace River in Alberta (Burgess<br />

1977a). P. strio/ata has recently gained in relative importance as a pest in both Saskatchwan<br />

and Alberta, being more abundant than P. cruciferae in some Saskatchewan rape<br />

fields after 1976 (Burgess 1981a) and in some fields in central Alberta since 1979 (H.G.<br />

Philip 1981 personal communication).<br />

Flea beetle problems in rape crops have increased in severity in the past decade. In<br />

Manitoba and Saskatchewan most fields <strong>of</strong> rape are threatened annually by these<br />

insects. In Alberta, flea beetle problems have tended to be less acute than in Saskatchewan<br />

and Manitoba, but in 1979 and 1980 heavy infestations were reported from<br />

most rape-growing areas (H.G. Philip 1981 personal communication). The estimated<br />

percentages <strong>of</strong> rape acreage treated with insecticides in 1979 were 75% in Manitoba,<br />

59% in Saskatchewan, and 45% in Alberta (Turnock & Samborski 1980). Although yield<br />

loss due to flea beetle damage is difficult to estimate. yield reductions <strong>of</strong> 15-50% have<br />

occurred in untreated plots in insecticide trials (Turnock & Samborski 1980). The<br />

combination <strong>of</strong> flea beetles and poor conditions for germination that occurred in 1975<br />

and 1980 in Manitoba completely destroyed many fields <strong>of</strong> rape.<br />

This is the first biological control programme set up against flea beetles <strong>of</strong> the genus<br />

Phyllotreta. The programme was initiated because <strong>of</strong> the low level <strong>of</strong> natural control <strong>of</strong><br />

P. cruciferae and P. strio/ata and because these are introduced species. Studies in<br />

Manitoba since 1975 confirmed only three native parasitoids <strong>of</strong> rape-infesting flea<br />

beetles: a braconid. Microctonus \'ittatae Mues.; a mermithid; and an allantonematid.<br />

Howardu/a sp. M. viltatae parasitizes adults <strong>of</strong> both P. cruciferae and P. strio/ala. but<br />

the incidence <strong>of</strong> parasitism is usually less than 5%. Various factors limit the impact <strong>of</strong><br />

this braconid in Manitoba. including its short adult life and low oviposition rate. and<br />

poor temporal synchronization with the host's population in mid-summer. Neither the<br />

mermithid nor the allantonematid is sufficiently abundant to limit populations <strong>of</strong> the<br />

73


74 H. G. Wylie, W. J. Turnock and L. Burgess<br />

Releases and Recoveries<br />

Table 11<br />

rape-infesting species. Occasional predation <strong>of</strong> flea beetle adults by adults <strong>of</strong> Collops<br />

viltalllS Say (Coleoptera: Melyridae) and Geocoris bullalus (Say) (Hemiptera:<br />

Lygaeidae) and by larvae <strong>of</strong> Chrysopa sp. (Neuroptera: Chrysopidae) have been reported<br />

(Gerber & Osgood 1975; Burgess 1977b, 1980).<br />

In Europe, a large complex <strong>of</strong> flea beetle species infest rape and other crucifers.<br />

Usually, the most destructive species are Psylliodes chrysocepha/a (L.) and Phyllolrela<br />

undu/ala Kutsch. (Carl 1973, Carl & Sommer 1975). Neither species occurs in <strong>Canada</strong>;<br />

P. chrysocepha/a was recorded once in Newfoundland (Anon. 1953) but apparently did<br />

not become established. Neither P. slrio/ala (sometimes called P. vinala F. by European<br />

entomologists; see Barber 1947) nor P. cruci/erae is considered to be an important pest<br />

in Europe. P. strio/ala is usually more abundant than P. erueiferae (Carl & Lehmann<br />

1980).<br />

Studies in Austria, Germany, and Switzerland identified three species <strong>of</strong> euphorine<br />

braconids, three mermithid species, and an allantonematid, Howardu/a phyllolrelae<br />

Oldham, as parasitoids <strong>of</strong> Phyllolrela spp. Although H. phyllolrelae parasitized up to<br />

70% <strong>of</strong> the flea beetle adults and reduced oviposition by parasitized females, it had little<br />

or no effect on beetle survival (Carl & Sommer 1975). The impact <strong>of</strong> the mermithids was<br />

also negligible. One <strong>of</strong> the braconid species was uncommon and was not identified. A<br />

second species, Microclonus sp. Z (Sommer 1981) which parasitized less than 10% <strong>of</strong><br />

the flea beetles, was morphologically indistinguishable from M. villalae (C.c. Loan<br />

1981 personal communication). The third braconid species, Mieroetonus bie%r<br />

Wesm., parasitized up to 50% <strong>of</strong> the flea beetle adults in summer. Small numbers <strong>of</strong> this<br />

braconid that were imported in <strong>Canada</strong> in 1977, and tested in the laboratory at Winnipeg,<br />

parasitized adults <strong>of</strong> P. eruci/erae and P. slrio/ala and developed on both species.<br />

Therefore, M. bie%r was considered to be a suitable candidate species for importation<br />

and release.<br />

Parasitoid cocoons received from the European Laboratory <strong>of</strong> the Commonwealth<br />

Institute <strong>of</strong> Biological Control were reared at the Research Program Service, Central<br />

Experimental Farm, Ottawa. Those cocoons which produced hyperparasites <strong>of</strong><br />

the genus Mesoehorus were destroyed, and adults <strong>of</strong> Mieroetonus bie%r and M. vittalae<br />

were shipped by air express to Winnipeg.<br />

Details <strong>of</strong> releases <strong>of</strong> Mieroetonus bie%r Wesm. are shown in Table 11. Parasitoids<br />

were released in the open in experimental plots <strong>of</strong> rape, Brassica napus L., cv. Tower, at<br />

the Glenlea Field Station <strong>of</strong> the Agriculture <strong>Canada</strong> Research Station, about 20 km south<br />

<strong>of</strong> Winnipeg. With the exception <strong>of</strong> three small releases, one in July 1979 and two in July<br />

1980 when host numbers were low, all parasitoids were released in May, June, August,<br />

or September when the plots were heavily infested by P. eruei/erae and lightly infested<br />

by P. strio/ata. No insecticides were applied to the plots each year for several weeks<br />

before the parasitoids were released, and none were applied after releases began.<br />

Open releases and recoveries <strong>of</strong> parasitoids <strong>of</strong> Phyllolreta spp.<br />

Species and Province<br />

M icroctonus bie%r Wesm.<br />

Manitoba<br />

Year Origin<br />

1978 Austria<br />

1979 Austria<br />

1980 Austria<br />

Number<br />

182<br />

375<br />

653<br />

Year <strong>of</strong> Recovery


Table 12<br />

Phyllotreta spp.. 75<br />

Several attacks by the parasitoid females on adults <strong>of</strong> P. strio/ata were observed on rape<br />

plants in the plots during releases.<br />

Flea beetles from the release sites were collected in 1979 and 1980 and reared in the<br />

laboratory for evidence <strong>of</strong> parasitoid establishment. Numbers collected were 8502 P.<br />

eruciferae, 261 P. strio/ata, and smaller numbers <strong>of</strong> other rape-infesting beetles in June<br />

1979 from the 1978 release site; 3082 P. cruciferae and 674 P. siriolaia during July­<br />

October 1979 from the 1979 release site; and 5039 P. siriolaia and 418 Psylliodes<br />

punetlllata Melsh. in April 1980 from overwintering sites adjacent to the 1979 release<br />

site. No M. hieolor emerged from these beetles.<br />

Some adults <strong>of</strong> M. bicolor and M. sp Z were retained for laboratory studies (Table 12).<br />

These studies showed that females <strong>of</strong> ,H. bicolor parasitized more P. striolClla than P.<br />

cruciferae when equal numbers <strong>of</strong> the two hosts were present.<br />

Imported parasitoids <strong>of</strong> Phyllolrela spp. used for laboratory studies<br />

Species<br />

Evaluation <strong>of</strong> Control Attempts<br />

Recommendations<br />

Microetonus hieolor Wesm.<br />

Mieroelonus sp. Z<br />

Year Origin<br />

1977 Germany<br />

1978 Austria<br />

Switzerland<br />

1979 Austria<br />

1980 Austria<br />

1977 Germany<br />

1978 Switzerland<br />

1979 Austria<br />

1980 Austria<br />

Number<br />

Several factors have limited the probability <strong>of</strong> establishment. First, small numbers <strong>of</strong><br />

parasitoids were released. Second. scarcity <strong>of</strong>the preferred host. P. striolala. at Glenlea<br />

may have limited oviposition by the females after release. Third. dispersal <strong>of</strong> the beetles<br />

in autumn to overwintering sites and in spring in search <strong>of</strong> food, scattered the population<br />

<strong>of</strong> parasitized beetles and thereby reduced the probability <strong>of</strong> female parasitoids that<br />

emerged in the spring being fertilized. If M. hieolor has become established, the small<br />

numbers released and scattering <strong>of</strong> the population would minimize our chances <strong>of</strong><br />

detecting establishment until the incidence <strong>of</strong> parasitism has increased substantially.<br />

Larger colonies <strong>of</strong> M. bieolor (> 50() females per colony) should be released. Although<br />

large numbers <strong>of</strong> parasitized flea beetles have been collected by the CIBC. only small<br />

numbers <strong>of</strong> parasitoid adults have been reared and female colony size has never exceeded<br />

100. There are two reasons for the small parasitoid colonies. First. many <strong>of</strong> the<br />

immature parasitoids enter diapause and eventually die when the hosts are reared in<br />

cages in Europe. Second. parasitoid emergence from cocoons obtained in Europe has<br />

usually been less than 50%. The possibility that higher temperature and longer photoperiod<br />

prevent diapause should be investigated. as well as the factors responsible for<br />

poor survival <strong>of</strong> the parasitoid larvae and pupae in cocoons.<br />

The possibility <strong>of</strong> obtaining additional candidate parasitoid species for release in<br />

<strong>Canada</strong> should be investigated in discussions planned in 1981 between staff <strong>of</strong> the CIBC<br />

and scientists from the USSR.<br />

43<br />

131<br />

20<br />

26<br />

8<br />

9<br />

2<br />

13<br />

I


Pest Status<br />

Background<br />

Releases and Recoveries<br />

Chapter 22<br />

Tetranychus urticae Koch, Twospotted<br />

Spider mite (Acarina: Tetranychidae)<br />

R.J. McCLANAHAN<br />

The twospotted spider mite. Tetranychus urticae Koch, continues to be a major problem<br />

in greenhouse crops, particularly cucumbers, chrysanthemums, and roses. The floral<br />

crops are fairly well protected by granular applications <strong>of</strong> the systemic insecticide<br />

aldicarb, but cucumber growers are having a difficult time keeping mites under control.<br />

Resistance to dic<strong>of</strong>ol and Pentac® (Zoecon Industries Ltd.) has developed in some<br />

greenhouses. The effective material oxythioquinox was withdrawn from use on cucumbers<br />

in 1974.<br />

The predaceous mite Phytoseiulus persimilis Athias-Henriot has been the only biological<br />

control agent seriously considered for control <strong>of</strong> T. urticae in greenhouses,<br />

although other mite predators are effective in orchards. The European workers have had<br />

encouraging results with P. persimilis and mass rearing is carried out at a number <strong>of</strong><br />

government and commercial establishments. Russian sources report a production <strong>of</strong> 83<br />

million P. persimilis per year from one rearing unit (Askretkov & Tolmacheva 1980).<br />

Probably the most extensive use <strong>of</strong> P. persimilis on greenhouse crops is in the Netherlands<br />

where some 400 ha (60% <strong>of</strong> total production) <strong>of</strong> cucumbers is protected by the<br />

predator (Koppert 1980).<br />

Research on the interaction <strong>of</strong> prey mites and P.persimilis has mostly been reported<br />

from Europe. Methods <strong>of</strong> introduction have been worked out for tomatoes (French et<br />

al. 1976) and cucumbers (Gould 1977), while the optimum conditions <strong>of</strong> temperature and<br />

humidity were defined by Stenseth (1979b) in Norway. Many references pertain to the<br />

compatibility <strong>of</strong> various insecticides and fungicides and the use <strong>of</strong> P.persimilis (e.g.<br />

Jeppson et al. 1975), but these results were superseded by investigations using a strain <strong>of</strong><br />

P. persimiiis resistant to phosphate insecticides (Stenseth 1979a).<br />

At Harrow, laboratory studies proved that some acaricides were selective in favor <strong>of</strong><br />

the predator (McClanahan 1971). However, the best materials were experimental or not<br />

widely accepted by growers for twospotted mite control. Since then oxythioquinox has<br />

been withdrawn from use on greenhouse cucumbers, and the twospotted mites have<br />

developed varying degrees <strong>of</strong> resistance to acaricides (McClanahan 1980).<br />

It is difficult to detail releases <strong>of</strong> P. persimilis in <strong>Canada</strong>, since there are several sources,<br />

including commercial enterprises, and only part <strong>of</strong> the production is used in <strong>Canada</strong>. A<br />

colony has been maintained at the Harrow Research Station for many years, and limited<br />

numbers are available on request by research establishments or educational projects.<br />

Acknowledgements <strong>of</strong> such shipments usually record successful establishment. P.<br />

persimilis are reared and sold by Better Yield Insects in Ontario and by Applied Bio­<br />

Nomics Ltd. in British Columbia. The number <strong>of</strong> predators available at the release point<br />

is dependent on the number <strong>of</strong> prey mites in the shipment and on the interval between<br />

77


78 R. J. McClanahan<br />

Recommendations<br />

Literature Cited<br />

packaging and releasing. Thus the number actually released can be much different than<br />

the number shipped, but under favorable conditions reproduction is very rapid with a<br />

generation in 5 days.<br />

Four commercial greenhouses near Leamington were used in a demonstration experiment<br />

in 1980. Two P. persimilis were placed on each mite-infested cucumber plant in<br />

February. Control <strong>of</strong> T. urlicae was achieved and maintained until late May, but then a<br />

change in greenhouse environment favored the pest mite and the growers had to switch<br />

to chemical control. Probably the humidity was lowered when vents were open<br />

throughout the day and this prevented P. pers;milis eggs from hatching.<br />

Techniques <strong>of</strong> using P. pers;milis in Canadian greenhouses will have to be refined and<br />

adjusted to greenhouse conditions. Introduction <strong>of</strong> predators onto the fall crop <strong>of</strong><br />

cucumbers might avoid a resurgence <strong>of</strong> T. urticae.<br />

Research projects could include modification <strong>of</strong> the greenhouse environment by<br />

misting to produce favorable conditions for the predator, or spraying only a portion <strong>of</strong><br />

the cucumber plants with an acaricide while predators were active on the unsprayed<br />

portion.<br />

Rearing under artificial conditions, as carried out in Russia (Storozhkov & Kamacheeva<br />

1980) should be tried in <strong>Canada</strong> for increased mass production.<br />

Askretkov. A.V.; Tolmacheva, KP. (1980) A sovkhoz biolaboratory. Zoshchita Rostenii 2,36 (in Russian).<br />

French, N.; Parr, W.J.; Gould. H.J.; Williams, J.l.; Simmonds, S.P. (1976) Development <strong>of</strong> biological control methods for the control <strong>of</strong><br />

Tetranychus urticae on tomatoes using Phytoseiu/us persimilis. Annals <strong>of</strong> Applied Biology 83, In-189.<br />

Gould, H.l. (1977) Biological control <strong>of</strong> glasshouse whitefly and red spider mite on tomatoes and cucumbers in England and Wales 1975 -76.<br />

Plant Pathology 26. 57-60.<br />

Jeppson, L.R.; McMurtry. 1.A.; Mead, D.W.; Jesser. M.J.; Johnson, H.G. (1975) Toxicity <strong>of</strong> citrus pesticides 10 some predaceous<br />

phytoseiid mites. Journal 01 Economic Entomology 68, 707-710.<br />

Koppen, P.C. (1980) Biological control in glasshouses in the Netherlands. Proceedings <strong>of</strong> the International Symposium on Integrated<br />

Control in Agriculture and Forestry, Vienna 1979. p. 484.<br />

McClanahan, RJ. (1971) Selective acaricides for integrated control <strong>of</strong> Tetranychus urticae. International Organization for Biological<br />

Control. West Palearctic Regional Section. Proceedings <strong>of</strong> a Conference on Integrated Control in<br />

Glasshouses. pp. 13-22.<br />

McClanahan. R.J. (1980) Why has integrated control practice levelled <strong>of</strong>f in <strong>Canada</strong>" Bulletin SROP 19801111, 141-144.<br />

Stenseth, C. (1979a) Effect <strong>of</strong> fungicides and insecticides on a resistant strain <strong>of</strong> Phytostiulus persimilis Athias-Henriot (Acarina:<br />

Phytoseiidae). Forskning og Forsok i Landbruket 30,77-83.<br />

Stenseth, C. (1979b) Effeet <strong>of</strong> temperature and humidity on Ihe development <strong>of</strong> Phytoseiulus persimilis and its ability to regulate populations<br />

<strong>of</strong> Tetranychus urtkae (Acarina: Phyloseiidae, Telranychidae). Entomophaga 24, 311-318.<br />

Storozhkov, Yu.V.; Kaznacheeva, V.P. (1980) Rearing <strong>of</strong> Phytoseiu/us in hydroponic greenhouses. Zoshchita Rostenii 2,34 (in Russian).


Pest Status<br />

Background<br />

Chapter 23<br />

Thymelicus lineola (Ochsenheimer),<br />

European Skipper (Lepidoptera:<br />

Hesperiidae)<br />

J.N. McNEIL<br />

In the last review dealing with the European skipper. Thymelicus lineola (Ochs.), Arthur<br />

(1971) reported that the distribution <strong>of</strong> this insect in <strong>Canada</strong> was limited to Ontario,<br />

Quebec. New Brunswick. and Nova Scotia in the east. with one localized population at<br />

Terrace, British Columbia. At that time economic losses caused by T. Iilleola had been<br />

recorded in Grey and Hastings Counties. Ontario. The distribution <strong>of</strong> the skipper has<br />

increased considerably and is now found in Prince Edward Island (Thompson 1974).<br />

Newfoundland (Jackson 1978). and in Manitoba (Preston & Westwood 1981). Also the<br />

population in British Columbia has now spread into the Lake Shuswap area (J. Proctor.<br />

1982. personal communication).<br />

Important hay losses have been recorded in Ontario (K. Bereza & W. Riley, 1978,<br />

personal communication. Quebec (McNeil et al. 1975), and Prince Edward Island<br />

(Thompson 1977). establishing it as a major pest species <strong>of</strong>timothy in the more northerly<br />

areas <strong>of</strong> its distribution. While the skipper is present throughout most northeastern<br />

states (Arthur 1971) the only incidence where T. lineola has reached pest status occurred<br />

in northern Michigan (R.F. Ruppel, 1977. personal communication).<br />

Bacillus thurillgiellSis Berliner. has proven effective in controlling larval skipper populations<br />

(Arthur & Angus 1965. Arthur 1968. McNeil et al. 1977) and is now recommended<br />

by the Conseil des productions vegetales du Quebec (1 x 10 10 I.U./ha) if control measures<br />

are required. It not only <strong>of</strong>fers satisfactory foliage protection when compared with chemical<br />

insecticides (Thompson 1977. Letendre & McNeil 1980), but also reduces the possible<br />

impact on non-target species, such as bees.<br />

The discovery <strong>of</strong> a nuclear polyhedrosis virus (NPV) affecting a high percentage <strong>of</strong> larval<br />

populations around Normandin. Quebec (Smim<strong>of</strong>f 1974) stimulated research to evaluate its<br />

potential as a biological control agent for T. lineola. Smim<strong>of</strong>f et al. (1976) and Duchesne<br />

(1980) demonstrated the efficacy <strong>of</strong> this virus. however its action is much slower than B.<br />

thurillgiellSis. Following an aerial application <strong>of</strong> B. thllringiellSis 1 x 10 10 I.U. at 18.71<br />

l/ha. infected larvae were observed within 2 days and 100% mortality was obtained in 6<br />

days (Fig. 2a). Defoliation levels at the time <strong>of</strong> harvest were 20%. the same a'i pre-treatment<br />

levels. while in the controls defoliation averaged 90%. By comparison an aerial application<br />

<strong>of</strong> the NPV. 5 x 10 6 polyhedra/ml at 18.71 l/ha. gave the same mortality but required<br />

considerably more time (Fig. 2b). Also defoliation at harvest was 50%, and while being<br />

lower than the 100% in the controls, was significantly higher than the pre-treatment level <strong>of</strong><br />

20%. Thus the virus has less potential than B. thllrillgiensis as a curative means <strong>of</strong> control,<br />

but could be introduced as a preventative measure if populations are beginning to increase<br />

but have not reached very high larval densities. This would be advantageous as the virus<br />

persists in the population and has an effect over several years (Duchesne 1980). While few<br />

other pathogens have been found infecting skipper larvae in Quebec. two entomopathogenic<br />

fungi ZoophtJlOra radicans and Entomophthora egressa have been isolated from larvae<br />

collected at Amqui in 1977 and 1978 (McNeil & MacLeod 1982).<br />

79


Imported Parasitoids<br />

Thyme/iells /it/('o/a (Ochsenheimer), 81<br />

Sixteen species <strong>of</strong> parasitoids have been reared from skipper populations in Quebec<br />

since 1972 and only 3 species are common to the 22 species found in Ontario (Arthur<br />

1962). As reported for Ontario total parasitism levels are very low and the only parasitoid<br />

recovered consistently in any numbers is Itoplectis cotlquisitor (Say). Based on<br />

observations over the last 7 years the importance <strong>of</strong> indigenous parasitoids in the<br />

population dynamics <strong>of</strong> T. lin eo/a would have to be considered as negligible. While the<br />

diversity <strong>of</strong> parasitoid species was lower in Europe, Carl (1968) reported high levels <strong>of</strong><br />

parasitism and felt that the principal larval parasitoids Pllryxe vulgaris Fall. and Rogas<br />

tristis Westm., as well as the major pupal parasitoid Syspasis (=Slenichneumon)<br />

scutellator (Grav.) would be suitable candidates for introduction into <strong>Canada</strong>. He<br />

believed that S. scutellator would be the most promising due to the fact that it was well<br />

synchronised with the skipper and, as it is univoltine, would not require alternate hosts<br />

as do the larval parasitoids.<br />

Specimens <strong>of</strong> P. vulgaris have been received regularly from Europe since 1972 (Table<br />

13), but unfortunately shipments usually arrived when local skipper populations had<br />

already pupated. Thus efforts were made to maintain laboratory cultures using Arlogeia<br />

rapae (L.) larvae as an alternate host. Female parasitoids readily oviposted and parasitoid<br />

larvae developed successfully, however difficulties in maintaining healthy host<br />

colonies inhibited successful mass rearing <strong>of</strong> the parasitoid. In 1974, adults arrived soon<br />

enough to be released directly in the field at Normandin (Table 14), but no P. vulgaris<br />

have been recovered in the area. Releases <strong>of</strong> adults and parasitized A. rapae larvae were<br />

made at Rawdon in 1974 and 1975 (Table 14), but were too late to coincide with the<br />

presence <strong>of</strong> skipper larvae. However cabbages were quite widely cultivated around the<br />

release site and an abundance <strong>of</strong> imported cabbageworm larvae were available as<br />

alternate hosts. In subsequent sampling no skipper larvae parasitized by P. vulgaris were<br />

recovered. It is <strong>of</strong> interest to note that P. vulgaris is an indigenous species and has been<br />

reared from A. rapae in Ontario (Harcourt 1966), yet has never been reported attacking<br />

T. lineola in either Ontario or Quebec. Carl (1968) suggested that different geographic<br />

races <strong>of</strong> P. vulgaris, with restricted host ranges may exist, which would explain this<br />

situation.<br />

Shipments <strong>of</strong> S. scutellator have been received from Delemont since 1974 and as seen<br />

in Tables 13 and 14, numbers were rather low in 1974 and 1975 but increased subsequently.<br />

In an effort to increase the numbers collected, Dr Carl treated fifth-instar skipper larvae<br />

with fluorescent powder and released them at the collection site at Plancher-Bas. It was<br />

felt that as the last larval skin remains attached to the pupae individuals could be easily<br />

located using a black light. While the recovery <strong>of</strong> marked individuals was poor, it was<br />

observed that skipper pupae that had been parasitized were more readily located with a<br />

black light than unparasitized ones, even without marking. Using this technique, night<br />

time collections yielded higher numbers than in previous years. Adults held under<br />

laboratory conditions throughout the winter (Table 13) were kept at 2 ± 1°C in plastic<br />

boxes containing leaf litter and peat moss. Males died rapidly and although females lived<br />

much longer only one survived until spring but died before T. lineola pupae were<br />

available. From 1977 all adults not used in open releases were held on arrival in an<br />

insectary and supplied with a 10% sugar solution as well as water. By mid-September all<br />

males and some females had died. At this time all surviving females were placed in large<br />

field cages containing logs, hay bales, and leaf litter. In 1977-78 there were no survivors<br />

while in 1978-79 we recovered 5 <strong>of</strong> the 200 females in mid-June. These were fed and<br />

held in the insectary until skipper pupae were available. One female stung the first pupa<br />

that she encountered, remaining 45 minutes with her ovipositor in the host and upon<br />

withdrawal, died. The stung pupa was held but no parasitoid emerged. The other 4<br />

females however, showed no interest in skipper pupae. New hosts were provided each<br />

day until the parasitoids died. All exposed pupae were held but only T. lineola adults<br />

emerged. In 1979 the holding cages were destroyed during a storm and most parasitoids


Blank Page<br />

84


Pest Status<br />

Background<br />

Chapter 24<br />

Tipula paludosa Meigen, European<br />

Cranefly (Diptera: Tipulidae)<br />

A.T.S. WILKINSON<br />

The early history <strong>of</strong> the European crane fly, Tipula paludosa Meigen, in <strong>Canada</strong> has<br />

been recorded already (Fox 1957. Wilkinson & MacCarthy 1967. Wilkinson 1971). In<br />

Nova Scotia, this pest peaked about 1965 (Creelman 1971) but had declined to a low level<br />

by 1971. High populations <strong>of</strong> the larvae (Ieatherjackets) damaged golf courses and lawns<br />

around St. John's, Newfoundland until 1970 (Creelman 1969, 1970). In the lower Fraser<br />

Valley <strong>of</strong> British Columbia, it has spread as far east as Hope. On Vancouver Island, it<br />

was found near Duncan and Victoria in 1972 and caused damage to golf courses and<br />

lawns in Victoria in 1976. It damaged lawns in Nanaimo in 1978 and in Alberni in 1979. In<br />

1980. it damaged lawns and a golf course in Prince Rupert (Fig. 3). Populations were<br />

highest when the pest was first found in 1965, but as the infestation was spread by gravid<br />

females from the introduction sites near Vancouver into the lower Fraser Valley and<br />

south into Washington State, populations declined, damage became less, and chemical<br />

controls became unnecessary. When new isolated infestations such as those on Vancouver<br />

Island and at Prince Rupert occur, they behave like those near Vancouver did<br />

and require control measures, at least in the first few years. The spread south in<br />

Washington was recorded by Jackson & Campbell in 1975, when crane flies had been<br />

found 180 km south <strong>of</strong> the British Columbia-Washington border.<br />

No indigenous parasitoids have been found in larvae <strong>of</strong> T. paludosa but many predaceous<br />

ground beetles and birds have been observed feeding on them. The starling<br />

continues to be the best predator <strong>of</strong> both the larvae and adults. Many birds feed on the<br />

adults.<br />

One disease, or combinations <strong>of</strong> the many diseases found in the larvae, may have been<br />

responsible for reducing populations and keeping them below the economic level, but<br />

this has not been proven. The following diseases have been identified in T. paludosa<br />

larvae from the Vancouver area by Dr. P.L. Sherlock, Rothamsted Experimental<br />

Station, Harpenden, Herts., England: Gregarina longa Leger, Hirmocystis venlricosa<br />

(Leger) Labbe, AClinocephaius lipuiae (Hammerschmidt) Leger, DiplocySlis sp. and<br />

Nosema binucleatum Weissenberg. All except N. binucleatum were also found in<br />

larvae from Newfoundland. In British Columbia, up to 50% <strong>of</strong> the larvae examined<br />

harbour all protozoans mentioned.<br />

The literature indicated that the parasitoid Siphona geniculata De Geer was considered<br />

the primary control agent <strong>of</strong> T. paludosa. In England, Rennie & Sutherland (1920)<br />

recorded 21.3% and IS.9% parastism for 2 consecutive years for the first generation and<br />

27.7% parasitism for the second generation. The latter was based on a low host population.<br />

Parasitism by S. geniciliala in other parts <strong>of</strong> Europe was not as high as that reported by<br />

Rennie & Sutherland. The Commonwealth Institute <strong>of</strong> Biological Control was retained<br />

by Agriculture <strong>Canada</strong>, Research Branch, in 1967 to study natural controls in Germany<br />

where T. paludosa occasionally causes damage in pastures. Populations <strong>of</strong> T. paludosa<br />

examined for parasitoids in 3 sites in Germany by Carl (1972) showed parasitism <strong>of</strong> 15.5,<br />

O.S, and 6.0% for the first generation and 4.9, 4.5, and 4.0% for the second generation. S.<br />

genicuiala has been recorded in Europe from Mamestra brassicae (L.), T. oleracea L., T.<br />

85


Releases and Recoveries<br />

Tipilia pailldosa Mcigen. 87<br />

fllivipennis De Geer. T. viltala Meigen. T. sllbnodicornis Zetterstedt. T. monlillm Egger.<br />

and T.maxima Poda (Rennie & Sutherland 1920. Chiswell 1956, Coulson 1962, Alma<br />

1975) as well as from T. pailldosa. It has been reared from only T. pailldosa in British<br />

Columbia.<br />

A phorid, Megaselia pailldosa (Wood), was found in T. pa/udosa by Coggins (1970) and<br />

by Carl (1972). It is rare in Europe and requires damp conditions.<br />

Many insecticides tested for the control <strong>of</strong> leatherjackets were effective. but only<br />

diazinon. chlorfenvinphos, and parathion were registered. Diazinon is used primarily on<br />

home and public lawns and the other two are used on pastures. Except in the more recent<br />

infestations at Nanaimo, Pon Alberni, Prince Rupen, and Victoria. there is very little<br />

need for chemical control. If these infestations follow the pattern <strong>of</strong> those in the lower<br />

Fraser Valley. populations will decline and chemical controls will no longer be needed.<br />

The B-exotoxin <strong>of</strong> Bacillus thuringiensis Berliner and the DD-136 nematode <strong>of</strong><br />

Neoplectana carpocapsae Weiser were investigated by Lam & Webster (1972) as possible<br />

controls for T. pailldosa. Both B-exotoxin and DD-136 were successful in killing<br />

leatherjackets in the laboratory, but the experimental evidence suggests that the heavy<br />

application required to achieve high mortality would not be practical for control <strong>of</strong> T.<br />

paludosa in the field.<br />

Carter (1973a. 1973b) showed that Tipula iridescent virus (TIV) can be transmitted<br />

when healthy larvae feed on a TIV-infected larva. Field trials by Caner (1978) showed<br />

that TIV can be introduced into field populations <strong>of</strong> T. paludosa but with low efficiency.<br />

In 1968. 331 adult S. geniculata were received from the Commonwealth Institute <strong>of</strong><br />

Biological Control (CIBC). Switzerland. from which 565 adults were reared and released<br />

near Cloverdale. British Columbia. Between 1971 and 1975 an additional 7000 S.<br />

geniculata received from the CIBC or reared at the Vancouver Research Station were<br />

released in the lower Fraser Valley. A technique for breeding S. geniculata was developed<br />

by Carl (1972). Leatherjackets were inoculated with larvae dissected from fieldcollected<br />

or laboratory-reared gravid females <strong>of</strong> S. geniculata. The first recovery was in<br />

1972 when 4 flies were reared from puparia attached to a leatherjacket cadaver. Counts<br />

<strong>of</strong> leatherjackets and parasitism were taken from January to August at Cloverdale from<br />

1974 to 1977 and at Ladner from 1975 to 1979. Parasitism was never very high. At<br />

Cloverdale. the highest parasitism recorded for the first generation (December to<br />

March) was 4.3% with an average <strong>of</strong> 1.6%, and for the second generation (June and July)<br />

the highest parasitism recorded was 5.1 % in 1974 and for the 4 years parasitism averaged<br />

2.4%. At Ladner during 5 years the highest parasitism was 3.3% for the winter generation<br />

with an average <strong>of</strong>2.3%. and for the second generation the highest parasitism recorded was<br />

6% with an average <strong>of</strong> 2.8%. As in Europe. the percent parasitism determined for the<br />

second generation was based on low numbers <strong>of</strong> the host because <strong>of</strong> the marked<br />

decrease in the larval population late in the season.<br />

In 1973. 100 S. geniculata adults were air expressed to Newfoundland; 79 survived<br />

and were released. In 1974. 600 were sent and released. None has yet been recovered.<br />

Megaselia pa/udosa was found in some <strong>of</strong> the leatherjackets received at the Vancouver<br />

laboratory from CIBC. Switzerland. It was reared for several generations but the<br />

colony died and the parasitoid was never released.<br />

Tipll/a iridescent virus-infected larvae were released near Ladner in 1975 but the virus<br />

has not been found in any larvae examined from this area.


Pest Status<br />

Background<br />

Chapter 25<br />

Trialeurodes vaporariorum (Westwood),<br />

Greenhouse Whitefly (Homoptera:<br />

Aleyrodidae)<br />

R.J. McCLANAHAN<br />

The greenhouse whitefly, Trialeurodes vaporariorum (Westwood), can be a problem<br />

wherever greenhouse vegetable and flower crops are grown. In major greenhouse areas<br />

<strong>of</strong> <strong>Canada</strong>, the damage has been alleviated considerably by extensive use <strong>of</strong> the<br />

parasitoid Encarsia formosa Gahan. When whiteflies are not controlled they cause<br />

reduction in plant vigour, formation <strong>of</strong> sooty mould on leaves and fruit, and reduced<br />

yield. Insecticide sprays, particularly the synthetic pyrethroids, are effective against the<br />

adults but repeated applications are required to keep the population at sub-economic<br />

levels.<br />

Much <strong>of</strong> the increased acreage under plastic and glass in the 1970s has been used for<br />

flower crops and bedding plants. These products may be infested with whitefly when<br />

they leave the greenhouse, in which case they have a lower value as well as serving to<br />

spread whitefly into new areas. Greenhouse whiteflies are also a problem in small hobby<br />

greenhouses or solar rooms <strong>of</strong> houses in which plants are grown. Biological control is<br />

the best means <strong>of</strong> combatting the pest in these situations.<br />

A research project on control <strong>of</strong> whiteflies was conducted at the Harrow Research<br />

Station from 1966 to 1974. An integrated control schedule was devised in which E.<br />

formosa were released, then sprays <strong>of</strong> oxythioquinox were applied to keep whiteflies<br />

under control until the parasitoids were established (McClanahan 1970, 1972). The<br />

withdrawal <strong>of</strong> oxythioquinox in 1974 shifted the emphasis to biological control alone. By<br />

this time mass rearing was well established and sufficient parasitoids were available to<br />

provide good whitefly control on greenhouse tomatoes and cucumbers (McClanahan<br />

1973).<br />

In Europe, renewed interest in biological control <strong>of</strong> whiteflies was generated by a<br />

need for control methods compatible with the use <strong>of</strong> Phytoseiulus persimilis Athias­<br />

Henriot for twospotted mite control. Extensive research was conducted in England and<br />

Holland on the interaction between E. formosa and whiteflies. A good review <strong>of</strong> this<br />

European research was given by Vet et al. (1980). Although there has been interest<br />

expressed in finding a whitefly parasitoid more effective than E. formosa at low<br />

temperatures, none has been found to date.<br />

The greenhouse whitefly was introduced to Japan about 1974 and within 4 years it<br />

became an important pest <strong>of</strong> greenhouse crops. A number <strong>of</strong> native parasitoids were<br />

found, not including E. formosa (Nakazawa & Hayashi 1977). Biological control is being<br />

considered in Japan.<br />

89


90 R. J. McClanahan<br />

Releases and Recoveries<br />

Table 15<br />

Evaluation <strong>of</strong> Control Attempts<br />

The use <strong>of</strong> E. fonnosa for whitefly control is so extensive that it is difficult to estimate<br />

the numbers released (Table 15). In the early 19705 production and distribution were<br />

handled by the Harrow Research Station. Parasitized whiteflies on tobacco leaves were<br />

packaged in lots estimated to contain 2, 5, or 10 thousand.<br />

A number <strong>of</strong> small businesses have been involved in E. fonnosa production and sales.<br />

Greenhouse Bio-Controls (1972-77) mainly supplied hobby greenhouse owners, with<br />

most sales to points in the United States. In 1973 and 1974 Greenhouse Bio-Controls<br />

supplied some <strong>of</strong> the needs <strong>of</strong> the Ontario Greenhouse Vegetable Producers' Marketing<br />

Board (OGVPMB) members, with 650000 and 947 000 parasitoids sold to the Board and<br />

distributed by them. This accounts for the increased Canadian distribution from this<br />

source in those years. In 1974-75, OGVPMB contracted E. formosa production to a<br />

grower, but that was not as successful as a subsequent arrangement whereby they<br />

provided partial salary for a technician supervised by Ontario Ministry <strong>of</strong> Agriculture<br />

and Food, with the rearing done in an isolated greenhouse at the Harrow Research<br />

Station.<br />

Releases <strong>of</strong> Encarsia formosa Gahan in <strong>Canada</strong> against Trialeurodes vaporariorum<br />

(Westwood).<br />

Source Numbers released- (thousands)<br />

1969 1970 1971 1972 1973 1974 1975 1976 1977 1978 1979 1980<br />

Harrow Research Station 10 150 943 3133 2148 1485 1157 6 IS 12 19<br />

Greenhouse BiD-Controls 14 886 1028 47 41 81<br />

OGVPMB" 500 750 2065 2160 lOSS 1265 550<br />

Beller Yield Insects 2S 40<br />

Applied Bio-Nomics 450 1750 588<br />

* Parasitized whitefly pupae from which there is 90-98% emergence <strong>of</strong> adult parasitoids.<br />

** Ontario Greenhouse Vegetable Producers' Marketing Board.<br />

In 1979 a similar arrangement was set up in British Columbia so growers there and in<br />

Alberta are supplied with E. formosa from Applied Bio-Nomics Ltd. The rearing<br />

facilities were supplied by Agriculture <strong>Canada</strong> and technical support was given by the<br />

British Columbia Department <strong>of</strong> Agriculture.<br />

Better Yield Insects was recently established in Ontario to produce and sell E.<br />

fonnosa and other biological control agents_ This source serves mainly hobby greenhouse<br />

owners in the United States, but plans were made to supply greenhouse vegetable<br />

growers in Ontario through OGVPMB, in 1981.<br />

Most releases have resulted in establishment, as judged by the very few failures<br />

reported. There have been cases where parasitoids were eliminated by an insecticide<br />

application, but with just one spray they have survived in the pupal stage and continued<br />

to multiply on the surviving whiteflies.<br />

The overall incidence <strong>of</strong> whitefly in Essex County has been reduced to the point where it<br />

is no longer a limiting factor in greenhouse tomato and cucumber production, and it is<br />

seldom a problem on field vegetables. E. formosa are easily found outdoors every<br />

summer.


Blank Page<br />

92


PART II<br />

BIOLOGICAL CONTROL OF<br />

WEEDS IN CANADA<br />

1969-80


Blank Page<br />

94


Success <strong>of</strong> the Programme<br />

Chapter 26<br />

Current Approaches to Biological Control<br />

<strong>of</strong> Weeds<br />

P. HARRIS<br />

A total <strong>of</strong> 30 agents has been released against 14 weed species in <strong>Canada</strong>. A third <strong>of</strong><br />

the agents have failed to become established, a third are established but inflict<br />

relatively minor damage to the weed, and a third inflict major damage. This gives a<br />

crude measure <strong>of</strong> the success <strong>of</strong> the programme, but it is not adequate to determine<br />

whether investment in biological control <strong>of</strong> weeds is justified in economic terms. The<br />

opportunity is taken in this chapter to discuss this and other current issues.<br />

The success <strong>of</strong> any pest control programme depends on a combination <strong>of</strong> its<br />

effectiveness and its cost. In biological control both <strong>of</strong> these vary greatly with the<br />

agent selected. It is easiest to predict the cost <strong>of</strong> an agent; so other things equal, the<br />

least expensive agent is the best one. Indeed, van Lenteren (1980) suggested that there<br />

are too many variables and unknowns to enable an effective agent to be selected on a<br />

scientific basis. Scientific or not, there are rules <strong>of</strong> thumb for increasing the chances <strong>of</strong><br />

success: matching the climate <strong>of</strong> the collection and release area (Wapshere 1980),<br />

collecting from the host plant species or strain on which the agent is to be released, and<br />

Harris (1973) suggested a number <strong>of</strong> desirable agent characteristics. These approaches<br />

can now be tested and modified against the record in Julien's (1982) world catalogue <strong>of</strong><br />

releases. The procedures that work in practice can then be followed, even if the<br />

reasons for success are not understood.<br />

Biological control <strong>of</strong> weeds, in common with government sponsored chemical weed<br />

control programmes such as those against leafy spurge and knapweed in <strong>Canada</strong>, has<br />

been derelict in not evaluating the benefits <strong>of</strong> the programme in economic terms.<br />

Without economic cost:benefit data it is difficult to determine whether a particular<br />

strategy against a weed makes economic sense or it is being pursued for political,<br />

philosophical, or historical reasons. Several chapters in this section represent initial<br />

evaluation <strong>of</strong> the prospective cost:benefits <strong>of</strong> biological control <strong>of</strong> possible target<br />

weeds. Hopefully, this is a step towards substantiating the intuition that biological<br />

control is economically superior to any other method <strong>of</strong> control for certain types <strong>of</strong><br />

weed problems.<br />

The usual practice in biological control is to rate individual agents as failures, partial<br />

or complete successes depending on their impact on the density <strong>of</strong> the host (see Laing<br />

& Hamai 1976). This is neat, but a problem, because it is sometimes the combined<br />

cropping pressure <strong>of</strong> several agents on a weed that achieves its density reduction; also<br />

the effects on weed density are <strong>of</strong>ten delayed by the seed accumulated in the soil so<br />

that the project cannot be rated until several years after the agent has achieved its<br />

maximum effect. I suggest that it is more useful to rate biological control agents <strong>of</strong><br />

weeds in terms <strong>of</strong> the proportion <strong>of</strong> the annual production removed or destroyed. The<br />

agents are similar in effect to cows in a pasture: if consumption is below a certain level,<br />

future yield is not reduced. Thus the good farmer restrains utilization <strong>of</strong> grasses to<br />

40-60% <strong>of</strong> annual production depending on the species and area. The good biological<br />

control worker, on the other hand, must try to exceed the threshold for the target weed<br />

(which might be determined by clipping tests). Most forbs are less tolerant than grasses<br />

to utilization (Jameson 1963), but many shrubs will tolerate browsing up to 75% <strong>of</strong><br />

95


96 P. Harris<br />

Table 16<br />

annual production (Ellison 1960). If the threshold <strong>of</strong> a target weed is assumed to be<br />

50% <strong>of</strong> annual production. it will not be controlled by an agent that removes 30% <strong>of</strong><br />

production. Nevertheless the addition <strong>of</strong> new defoliators, seed or root feeders, will<br />

normally increase the total amount <strong>of</strong> the weed consumed. Thus the practice is to add<br />

agent species until the threshold is exceeded. For this reason I consider any agent that<br />

has become numerous enough to remove a substantial proportion <strong>of</strong> annual production<br />

as a success. I would prefer to rate them by the actual amount removed but do not<br />

have the data for all the species established in <strong>Canada</strong>.<br />

The successes <strong>of</strong> the Canadian programme are listed in Table 16. Several <strong>of</strong> the<br />

Agents removing a substantial proportion <strong>of</strong> the annual production <strong>of</strong> the target weed in<br />

one or more regions <strong>of</strong> <strong>Canada</strong>.<br />

Target Weed<br />

Carduus nutans L.<br />

Centaurea diffusa Lam.<br />

e. maculosa Lam.<br />

Cirsium vulgare (Savi) Ten.<br />

Euphorbia cyparissias L.<br />

Hypericum per/oratum L.<br />

Senecio jacobaea L.<br />

Agent<br />

Rhinocyllus conicus Froel.<br />

Urophora affinis Frfld.<br />

U. quadri/asciata (Mg.)<br />

Urophora sty lata L.<br />

Hyles euphorbiae (L.)*<br />

Anaitis plagiala (L.)<br />

Chrysolina hyperici (Forst.)<br />

C. quadrigemina (Suffr.)<br />

Tyria jacobaeae L.<br />

Longitarsus jacobaeae (Wat.)<br />

• The inclusion <strong>of</strong> this insect is possibly an optimistic estimate <strong>of</strong> its effect.<br />

Country<br />

Screening Agent<br />

<strong>Canada</strong><br />

<strong>Canada</strong><br />

<strong>Canada</strong><br />

<strong>Canada</strong><br />

<strong>Canada</strong><br />

<strong>Canada</strong><br />

Australia<br />

Australia<br />

New Zealand<br />

United States<br />

agents have in fact achieved density reductions <strong>of</strong> the target weed that have solved<br />

some <strong>of</strong> the problems from it. For example, the seedhead weevil, Rhillocyllus conicus<br />

FroeI., has reduced the density <strong>of</strong> Carduus nutans L. sufficiently on Saskatchewan<br />

rangeland that it does not threaten the cattle stocking rate. On the other hand, the<br />

thistle remains numerous on disturbed sites such as gravel pits, and even at low<br />

densities can be a nuisance in parks and around beaches. Similarly Hypericum<br />

perforatum L. presently has little effect on the stocking rate <strong>of</strong> British Columbia<br />

rangeland: the dense stands at Elko, British Columbia, are not <strong>of</strong> concern for cattle<br />

production as stocking is kept low to retain the area as a wildlife overwintering refuge.<br />

The weed, however, is still perceived to be a problem by the ranchers, possibly<br />

because they fear its spread.<br />

The initial work on four <strong>of</strong> the successful agents (Table 16) was done by other<br />

countries. Insects successful in Australia or California, such as Chrysolina quadrigemina<br />

(Suffr.), have tended to be successful in <strong>Canada</strong> despite differences in climate and the<br />

converse is true <strong>of</strong> Canadian pioneered species such as R. conicus. Indeed, the most<br />

effective and widely employed technique for selection <strong>of</strong> an effective agent is to choose a<br />

proven winner. Such agents are also inexpensive as most <strong>of</strong> the pre-release studies<br />

required by <strong>Canada</strong> have been done. Stock for the Chrysolina spp. established in<br />

<strong>Canada</strong> was obtained from California following success there. It would almost certainly<br />

have achieved an impact on the weed in <strong>Canada</strong> sooner if the releases had been made<br />

with climatically pre-adapted stock; but there is a trade-<strong>of</strong>f between speed <strong>of</strong> success<br />

and cost. In North America, savings can be achieved on weeds common to both <strong>Canada</strong>


Table 17<br />

Current approaches to biological control <strong>of</strong> weeds 97<br />

and the United States by demonstrating the effectiveness <strong>of</strong> the agent in one country and<br />

only distributing those that are successful.<br />

There are 11 agent species established in <strong>Canada</strong> that are not removing a substantial<br />

proportion <strong>of</strong> the host plant production (Table 17). The reasons for their failure to<br />

increase to a high population density vary. Some were released so recently that there has<br />

not been time. Sphenoptera jugosfavica Zell. which was released in 1977 has increased<br />

to a high density within 0.25 km <strong>of</strong> the release point but is still too restricted in<br />

distribution to be included in Table 16. Other species may require a period <strong>of</strong> climatic<br />

adaption through selection. Studies on the successful agents Tyria jacobaeae L., Hyles<br />

euphorbiae (L.), and Chrysolina quudrigemina indicated that a period <strong>of</strong> 2-13 generations<br />

<strong>of</strong> natural selection was required to produce a population adapted to the release<br />

area. Thus several <strong>of</strong> the insects in Table 17 are possible successes.<br />

Agents established in <strong>Canada</strong> but not removing a substantial proportion <strong>of</strong> annual<br />

production <strong>of</strong> target weed.<br />

Target Weed<br />

Carduus nutans L.<br />

Centaurea diffusa Lam.<br />

C. maculosa Lam.<br />

Acroptilon repens (L.)<br />

Cirsium arvense Scop.<br />

Euphorbia esula-virgata<br />

complex<br />

Hypericum perforatum L.<br />

Linaria vulgaris Mill.<br />

Sonchus arvensis L.<br />

Carduus acanthoides L.<br />

Agent<br />

Trichosirocalus ho"idus Panz.<br />

Sphenoptera jugoslavica Zell.<br />

Metzneria paucipunctella Zell.<br />

Paranguina picridis Kirj. & Ivan.<br />

Ceutorhynchus litura F.<br />

Urophora cardui (L.)<br />

Oberea erythrocephala Schr.<br />

Aphis chloris (Koch)<br />

Calophasia lunula (Hufn.)<br />

Cystiphora sonchi (Bremi)<br />

Rhinocyllus conicus Froel.<br />

Year Established<br />

1978<br />

1977<br />

1974<br />

1977<br />

1968<br />

1976<br />

1981<br />

1981<br />

1966<br />

1982<br />

The failure <strong>of</strong> R. conicus to attack a high proportion <strong>of</strong> Carduus acanthoides L. flower<br />

heads seems to be related to a poor adaptation to this thistle. The weevil was a success<br />

on C. nutans in <strong>Canada</strong>, the host plant from which it was collected in Europe. The<br />

association <strong>of</strong> both specialized insects and pathogens with a host plant is an extremely<br />

stable relationship. The completion <strong>of</strong> a number <strong>of</strong> generations on a marginal host<br />

normally does not improve survival unless the genetic potential to overcome host<br />

resistance already exists in the agent population. There are host races <strong>of</strong> R. conicllS on a<br />

number <strong>of</strong> European thistles (Zw6lfer & Preiss 1983); but it is uncommon on C.<br />

acanthoides and was not found in the samples examined by Zw6\fer (1965). Thus the C.<br />

nutans strain <strong>of</strong> R. conicus is likely to remain marginal on C. acanthoides as the species<br />

does not have the genetic potential to exploit it successfully.<br />

Two <strong>of</strong> the 12 failures listed in Table 18 have been established in <strong>Canada</strong> on the plant<br />

from which they were collected in Europe. Thus only ten or a third <strong>of</strong> the agents have<br />

failed to become established at all. This compares with a failure rate <strong>of</strong> about two thirds


Number <strong>of</strong> Phytophages and Weed Control<br />

Current approaches to biological control <strong>of</strong> weeds 99<br />

depressed by overgrazing or disturbance. Canadian examples are H. perfora/um,<br />

Centaurea diffusa Lam., C. maculosa Lam. in the grasslands <strong>of</strong> the interior <strong>of</strong> British<br />

Columbia; Carduus nlllans in the mid-grass prairie <strong>of</strong> Saskatchewan; Euphorbia esulavirgala<br />

complex in Manitoba; and E. cyparissias L. on limestone soils in<br />

Ontario, North American plants such as Solidago gigantea Ait, behave similarly in<br />

parts <strong>of</strong> Europe. These plants do not form extensive monocultures in their native region<br />

where they are cropped by phytophages. For example, E. cyparissias normally forms<br />

only a small portion <strong>of</strong> the herbaceous community on limestone soils in Europe compared<br />

to 25-50% at Braeside, Ontario. The introduction <strong>of</strong> phytophages against H.<br />

perforalum in British Columbia and Ontario reduced it to a small percentage <strong>of</strong> its<br />

former abundance at most sites, and C. nlllans has been reduced to less than 10% on<br />

Saskatchewan rangeland. In part, the effectiveness <strong>of</strong> a control agent depends on the<br />

level <strong>of</strong> competition from other plants (Harris 1981a) which is normally greatest in<br />

uncultivated habitats and least in disturbed sites.<br />

Specialized phytophages and their hosts normally have a mutual density dependence.<br />

Thus cropping pressure is high when the plant is abundant and low when it is scarce and<br />

vice versa for the mortality <strong>of</strong> the phytophage. The result is that neither becomes<br />

exterminated. A Canadian biological control example is the cinnabar moth, Tyria<br />

jacobaeae, on Senecio jacobaea L. in British Columbia in which the moth has<br />

tracked the density <strong>of</strong> its host (Lakhani & Dempster 1981). Similarly Ralph (1977) found<br />

that the highest densities <strong>of</strong> milkweed (Asclepias syriaca L.) supported the highest<br />

densities <strong>of</strong> Oncopellus fasciatus (Dall.), and that below a certain density that plant<br />

escaped attack. This has been called 'escape in space' by Feeney (1976).<br />

To summarize, classical biological control <strong>of</strong> weeds is most appropriate in terms <strong>of</strong> its<br />

impact and its effects against introduced weed species that dominate large areas <strong>of</strong> range<br />

or other uncultivated land.<br />

Cropping pressure on a weed can be increased by establishing several phytophage<br />

species on it (Harris 1981a). Even two closely competing seed head flies on C. diffusa<br />

reduced seed production more when together than either did alone (Myers & Harris<br />

1980). Internationally, introductions have tended to continue over a period <strong>of</strong> years until<br />

an average <strong>of</strong> four species has been established on the weed (Harris 1979). This process<br />

<strong>of</strong> adding agents can only go so far as there is a species area-asymptote: the number <strong>of</strong><br />

phytophagous insects and pathogens cropping a plant species varies with the log <strong>of</strong> its<br />

abundance (Lawton & Schroeder 1977, Strong & Levins 1979). That is to say the<br />

number <strong>of</strong> phytophages that can be supported by a plant species doubles for every<br />

tenfold increase in its abundance. In southern Britain five continental species <strong>of</strong> Heteroptera<br />

have become established on pine following large plantings. The pines displaced<br />

juniper and as a result one juniper-feeding Heteroptera, which formerly existed in<br />

sizeable colonies, has become extinct (Southwood 1957). Simberl<strong>of</strong>f (1978) achieved<br />

similar results experimentally by altering the size <strong>of</strong> mangrove islands. In some plants an<br />

expansion <strong>of</strong> range is matched by an increase in the species consuming it. The recruits<br />

are derived from other plants in the new range or from the host in the native habitat. This<br />

occurred within 50 years on cacao (Strong 1974). With other introduced plants such as E.<br />

cyparissias in North America, there has been negligible recruitment <strong>of</strong> insect consumers<br />

even after a hundred years. Presumably most native insects are unable to overcome the<br />

physical and chemical defenses <strong>of</strong> the plant and the infestations are too remote for<br />

specialized insects to reach them from Europe.


100 P. Harris<br />

The number <strong>of</strong> niches available to insect species is partly a function <strong>of</strong> the plant<br />

architecture, so a species native to several parts <strong>of</strong> the world tends to support similar<br />

numbers <strong>of</strong> leaf chewers, stem borers, and insects in other guilds throughout its range,<br />

even though the insects may be in different taxonomic orders (Lawton 1978). The niches<br />

can be regarded as sites at which the nutrient pool in the plant can be tapped. The pool<br />

normally cannot be drained through anyone tap but they all lower its level. This implies<br />

competition between agents in different niches. For example, Moran & Southwood<br />

(1982) found that there was interaction between chewers and sap suckers on the tree<br />

species studied.<br />

The best targets for classical biological control are introduced plants on which the<br />

consumer pressure is below that found in the native region. It should be possible to raise<br />

consumption to at least the native region level although the number <strong>of</strong> phytophages that<br />

can be established will be lower as their density will not be depressed by specialized<br />

parasites. Probably each agent established on a weed increases the difficulty <strong>of</strong> establishing<br />

additional species so those with the most potential should be introduced first.<br />

It is unlikely that native weeds can be controlled by classical biological control unless<br />

their abundance has increased. This has occurred with ragweed, Ambrosia artemis;;folia<br />

L. (Harris & Piper 1970). If the phytophage-ragweed area asymptote is below that<br />

<strong>of</strong> ragweeds that have not increased since European settlement, it should be possible to<br />

establish additional agents from South America; however the increased consumer<br />

pressure on the weed may not reduce allergenic hay-fever since the symptoms in man<br />

are relatively dose independent. Usually the best hope for the biological control <strong>of</strong> a<br />

native weed is the periodic application <strong>of</strong> an agent (probably a native pathogen) to keep it<br />

at an artificially high level. This is called augmentative biological control and is discussed<br />

under the heading 'The Future'. Hall et al. (1980) found no significant difference in<br />

the success <strong>of</strong> biological control against native and exotic insect pests, but whether the<br />

native pests had increased in adundance with modem agriculture was not considered.<br />

Demonstration that Candidate Agents are not a Threat to Desirable Plants<br />

The use <strong>of</strong> a plant species as a host by a phytophage depends on its suitability as a<br />

nutrient source and as a place to live, the opportunity <strong>of</strong> reaching the plant, and the<br />

survival advantage <strong>of</strong> using it (Harris 1981b). If one <strong>of</strong> these requirements is<br />

unfavourable, the agent will not reach high levels on it. The strategy is to show that the<br />

organism cannot develop on any plant outside a taxonomic group that does not contain<br />

desirable species (Harris & Zwolfer 1968, Zwolfer & Harris 1971, Wapshere 1974).<br />

Unfortunately in the laboratory, organisms develop on a broader range <strong>of</strong> plants than<br />

they attack in nature. It is then necessary to show that the organism either does not<br />

have the opportunity <strong>of</strong> establishing on the desirable plant in nature or that individuals<br />

doing so are at a selective disadvantage. In effect there is a 3 tier system <strong>of</strong> testing:<br />

desirable plants that fail the first test because they will support development, are<br />

considered at the next tier <strong>of</strong> opportunity and if this is positive, at the last, and most<br />

difficult, tier <strong>of</strong> advantage.<br />

The opportunity <strong>of</strong> an insect to attack a plant within its distributional range normally<br />

depends on the presence in the plant <strong>of</strong> sensory tokens for attraction and acceptance<br />

by the ovipositing female. This is the result <strong>of</strong> many generations <strong>of</strong> selection to<br />

optimize survival and is not easily changed. Indeed, it is <strong>of</strong>ten difficult to establish an<br />

agent if the strain or species <strong>of</strong> the target weed is different from the one on which it was<br />

collected. This is evident in the programme for the biological control <strong>of</strong> leafy spurge in<br />

<strong>Canada</strong>. Similarly Goeden (1978) found that the race <strong>of</strong> R. conicus established on C.<br />

nutans in North America did not attack Silybum marianum (L.) Gaertn. in the field,


Current approaches to biological control <strong>of</strong> weeds 101<br />

although a strain collected from this plant established readily. This programming <strong>of</strong> an<br />

insect for a particular plant tends to break down in the laboratory and so is best field<br />

tested in the native region.<br />

In contrast to the active dispersal <strong>of</strong> most insects, the passive dispersal <strong>of</strong> pathogens<br />

and to a slightly lesser extent certain insects like aphids, results in their deposition on<br />

many plant species which they will frequently attempt to attack. The host range <strong>of</strong><br />

pathogens is normally genetically stable so most <strong>of</strong> the pathogen propagules landing on<br />

other plant species are lost. However, if the genetic capacity (possibly as a result <strong>of</strong><br />

hybridization) to attack another plant species exists in a pathogen popUlation, the<br />

occasionally successful individual can, by asexual reproduction, increase rapidly. It is<br />

not a good omen for biological control if dense stands <strong>of</strong> a desirable plant on which a<br />

pathogen can develop in the laboratory, grow near the target weed.<br />

The advantage <strong>of</strong> a particular host species to a phytophage depends on its potential for<br />

increase on it. The innate reproductive capacity <strong>of</strong> the phytophage on various plants can<br />

be measured in the laboratory, but the rate <strong>of</strong> increase in the field is influenced by<br />

mortality, competition from other consumers, the density <strong>of</strong> the host, and other factors<br />

that cannot be measured in the laboratory. Thus a good laboratory host is not necessarily at<br />

risk in nature, and a relatively poor laboratory host might be regularly damaged if there<br />

are large amounts growing as a monoculture. This is hard to determine experimentally in<br />

the new habitat without danger that the organism will escape. For this reason the<br />

development <strong>of</strong> the organism on an introduced crop plant in the laboratory is justification<br />

for rejecting it as a biological control agent; but development on a native wild plant is less<br />

cause for concern. For example, the European moth Plryllonorycter blancardella<br />

(Fabr.) attacks Cratueglls spp., apple, and several related plants in its native range but,<br />

where it has been introduced into Quebec, Pottinger & LeRoux (1971) were unable to<br />

find it on Cratueglls spp., an abundant native, although it was common on the introduced<br />

apple. Similarly Moran (1982, personal communication) found that insect pests on South<br />

African crops derived from native plants were, with rare exceptions, attacked by native<br />

species. On the other hand, introduced crops were attacked by both native and introduced<br />

insects.<br />

There are examples in which introduced insects and pathogens have become pests <strong>of</strong><br />

native plants. These phytophages are either polyphagous, like the gypsy moth, so they<br />

readily accept and develop on a wide range <strong>of</strong> plants; or the native lacks resistance, as in<br />

the case <strong>of</strong> the chestnut blight, so the organism has a higher reproductive capacity on it<br />

than on the original host. The best indication <strong>of</strong> the genetic capability <strong>of</strong> phytophages to<br />

utilize plants in the new habitat is their host range in the native region, and if necessary<br />

desirable plants <strong>of</strong> the new habitat can be tested there by planting them at their normal<br />

density. Thus a species such as Lema cyan ella L., which in Europe is only found on<br />

Cirs;llm arvense (L.) Scop. (Peschken & Johnson 1979) is unlikely to attack native<br />

North American thistles to any extent even though it will accept them in the laboratory.<br />

Selection <strong>of</strong> Target Weeds for Classical Biological Control<br />

Classical biological control cannot escape some measure <strong>of</strong> politics because it depends<br />

on public funding. Also, as an agent will spread over many properties, it has to be<br />

introduced as a matter <strong>of</strong> public interest.<br />

Originally the priorities for biological control in <strong>Canada</strong> were largely determined by<br />

popular demand: the squeaky wheel approach. This did not necessarily target the best or<br />

the most urgent weeds for biological control as is evident from some <strong>of</strong> the examples in<br />

this volume.


102 P. Harris<br />

The Future<br />

Literature Cited<br />

It is axiomatic that biological control is only justified if it is likely to produce a<br />

satisfactory return on investment which is greater than can be obtained by other means<br />

<strong>of</strong> control. The biological control <strong>of</strong> a suitable target weed is likely to require 20 scientist<br />

years ($2 million at present costs) (Harris 1979). These costs are less if host specificity<br />

testing has been done elsewhere. Determination <strong>of</strong> benefit requires quantification <strong>of</strong><br />

savings and increased yields minus any detrimental effects <strong>of</strong> control such as the<br />

reduction <strong>of</strong> a nectar source. Costs and benefits that cannot be quantified in monetary<br />

units should be expressed in ecological terms. The study should be published before<br />

biological control is started. This gives an opportunity for objections to be voiced and<br />

the proposal debated.<br />

Classical biological control <strong>of</strong> weeds is likely to continue to be directed primarily against<br />

introduced species that form extensive dense stands on uncultivated land. For this type<br />

<strong>of</strong> problem there is a hgh rate <strong>of</strong> success and the return is excellent, particularly when<br />

considered on.a national or a continental basis. There are, however, a limited number <strong>of</strong><br />

weed species that are prime targets in <strong>Canada</strong>. As work on these is completed or the<br />

possibilities for biological control are exhausted, as is happening for C. arvense, the<br />

programme will be directed increasingly against weeds <strong>of</strong> more minor importance. One<br />

way <strong>of</strong> maintaining a favorable benefit-cost ratio is to tackle them as joint projects with<br />

the United States or other countries so that the costs are shared and the benefits accrue<br />

on an international scale. At present international cooperation is informal and this will<br />

probably need to be replaced by more formal cooperative agreements against specific<br />

target weeds.<br />

There is room to increase the rate at which present projects are completed but the<br />

long-term prospect is for a continued small Canadian participation in classicial biological<br />

control <strong>of</strong> weeds. There has been an increase in staffing since the 1968 review (Harris<br />

1971) that reflects the economic seriousness <strong>of</strong> some <strong>of</strong> the projects such as knapweed<br />

and leafy spurge. Thus Agricultural <strong>Canada</strong> has increased its participation from 2 to 4<br />

full-time scientists; McGill University (Macdonald College) has an active programme<br />

that emphasizes pathogens; the Province <strong>of</strong> Alberta is recruiting a scientist, and the<br />

University <strong>of</strong> British Columbia has a programme elucidating the effects <strong>of</strong> some <strong>of</strong> the<br />

agents. The programme will probably stay at this level with more emphasis placed on the<br />

selection <strong>of</strong> both the target weeds and the agents.<br />

The main area for expansion is in augmentative biological control in which a pathogen<br />

is applied periodically as a bioherbicide. Two centres, Macdonald College and the<br />

Regina Research Station have started investigations in this area. The work has not been<br />

reported in this review as it has yet to lead to the licensing <strong>of</strong> a bioherbicide in <strong>Canada</strong>.<br />

The few bioherbicides currently in use in the United States are for weeds <strong>of</strong> crops that<br />

are hard to control by other means (Templeton 1982). One <strong>of</strong>the attractions <strong>of</strong> bioherbicides<br />

is that they can be produced in small amounts at a reasonable cost. Thus they can<br />

be used for weed problems <strong>of</strong> minor crops in which the area involved does not justify the<br />

development <strong>of</strong> a specific herbicide. The method has a pr<strong>of</strong>it potential so it is likely to be<br />

developed primarily by industry and this could occur rapidly. The main obstacle to date<br />

has been the lack <strong>of</strong> licensing regulations. These have now been drafted in the United<br />

States and it is probable that <strong>Canada</strong> will adopt similar requirements. There is an initial<br />

role for government research to get the industry started. Subsequently the government<br />

role may be largely regulatory.<br />

Ellison, L. (1960) Influence <strong>of</strong> grazing on plant succession <strong>of</strong> range plants. Botanical Review 26, 1-66.<br />

Feeney, P. (1976) Plant apparency and chemical defense. In: Wallace, l.W.; Mansell, R.L. (Eds.). Biochemical interactions between plants<br />

and insects. Recent Advances in Phytochemistry 10, 1-40. New York; Plenum Press.


Blank Page<br />

104


Pest Status<br />

Background<br />

Chapter 27<br />

Acroptilon repens (L.) DC., Russian<br />

Knapweed (Compositae)<br />

A.K. WATSON and P. HARRIS<br />

Acrop,ilon repens (L.) DC. (Centaurea repens L.) is native to Mongolia. Western<br />

Turkestan, Iran, Turkish Armenia and Asia Minor (Moore & Frankton 1974). and was<br />

first introduced into <strong>Canada</strong> about 1900 as a contaminant <strong>of</strong> Turkestan alfalfa seed (Groh<br />

1940). Russian knapweed is now widespread in the southern parts <strong>of</strong> the four western<br />

provinces and is also found in southern Ontario. although most infestations are relatively<br />

small (50% <strong>of</strong> infestations in Saskatchewan are less than one acre in size) (Watson 1980).<br />

This noxious weed has a well developed, extensive root system which is the major<br />

means <strong>of</strong> reproduction and spread (Frazier 1944). It forms dense infestations in cultivated<br />

fields, in grain and alfalfa fields. in pastures, along roadsides and irrigation ditches and in<br />

waste places. Russian knapweed is apparently indifferent to crop association and is able<br />

to survive in almost any crop in tillable soil (Rogers 1928). Infested fields commonly<br />

have knapweed densities <strong>of</strong> 100-300 shootslm 2 which suppress and essentially eliminate<br />

other plant growth (Selleck 1964). In addition to its intense competitive ability, Russian<br />

knapweed is poisonous to horses (Younget al. 1970). Actual crop losses and other losses<br />

due to Russian knapweed have not been determined nor estimated in <strong>Canada</strong>.<br />

Because <strong>of</strong> its serious threat to Canadian agriculture, Russian knapweed was designated<br />

as a "prohibited noxious weed" in the Federal Seeds Act (Agriculture <strong>Canada</strong><br />

1967). The benefit <strong>of</strong> this is debatable as it is difficult to collect viable seed in <strong>Canada</strong><br />

(Watson 1975) and germination rarely occurs in the field (Selleck 1964). Control <strong>of</strong> this<br />

persistent perennial weed is difficult as its extensive root system is not adversely<br />

affected by cultivation and the weed is relatively resistant to commonly used herbicides.<br />

Picloram (4-amino-3,5.6-trichloropicolinic acid) and glyphosate [N-(phosphonomethyl)<br />

glycine) are two <strong>of</strong> the more effective herbicides for Russian knapweed control (Alley &<br />

Humberg 1979). The biology <strong>of</strong> Russian knapweed has recently been reviewed (Watson<br />

1980).<br />

In North America, Russian knapweed is relatively free <strong>of</strong> specialized parasites and is not<br />

extensively attacked by polyphagous feeders, but in its native range, Russian knapweed<br />

is the host <strong>of</strong> a number <strong>of</strong> specialized organisms (Watson 1980). The potential biological<br />

control agents <strong>of</strong> Russian knapweed include seven organisms that attack the seed head:<br />

a seed gall mite (Aceria acropliloni V. Shev. & Kov.), three Diptera (Dasyneura sp.,<br />

Urophora maura (Frfld.), Urophora kasochstanica V. Richter) and three Coleoptera<br />

(Larinus bardus Gyll., Larinus jaceae Fabr., Rhynchaenus dislans Faust); a stem gall<br />

former (Aulacida acroptilonica Beliz); a leaf and stem rust (Puccinia acroptili Syd.); and<br />

a leaf and stem gall nematode (Paranguina picridis Kirj. & Ivan.) (Watson 1980).<br />

105


108 A. K. Watson and P. Harris<br />

Table 21<br />

Table 22<br />

Recovery <strong>of</strong> P./Jicridis Kirj. & Ivan. on Russian knapweed, A. repem (L.) DC., at the<br />

Regina Research Station.<br />

Precipitation<br />

Year Galls/m: P. picridislm: A. repetls stems/m: September-May (mm)<br />

1977 0<br />

1978 0.67<br />

1979 3.0<br />

1980 no obvious<br />

galls but some<br />

swollen stems<br />

1981 0.63<br />

0<br />

2.9<br />

3599<br />

16<br />

15.0<br />

52.0<br />

18.6<br />

167.5<br />

212.5<br />

213.0<br />

145.7<br />

108.6<br />

The nematode population declined in the springs <strong>of</strong> 1980 and 1981, which was likely the<br />

result <strong>of</strong> the dry conditions. In the five years from release, no P. picridis were found in the<br />

control strip <strong>of</strong> knapweed planted 3 m away from the test plots.<br />

At the Ste-Anne-de-Bellevue site, 6 m l plots were established in the early spring <strong>of</strong><br />

1977 and 100 000 nematodeslm l were inoculated in three <strong>of</strong> the plots. Russian knapweed<br />

was transplanted into all plots and eight crop species were sown in rows between the<br />

Russian knapweed plants in 1977. In 1978. globe artichoke and safflower were sown in<br />

all plots. A few nematode galls were observed on Russian knapweed in 1977 and heavy<br />

galling occurred on Russian knapweed in 1978, 1979, and 1980. In 1981, gall formation on<br />

Russian knapweed was reduced. Galls did not develop on any <strong>of</strong> the tcst plant species in<br />

1977, 1978, or 1979.<br />

Effect <strong>of</strong> P. picridis Otl the ktlapweed<br />

P. picridis was innoculated into 3 m l plots <strong>of</strong> pure Russian knapweed at Jenner, Alberta.<br />

at a rate <strong>of</strong> 30 Ooo/ml, and three control plots were established with a 50 em buffer zone<br />

from the treated area.<br />

No galls were found on the knapweed in the 1978 growing season even though the<br />

nematodes had been kept in soil for six weeks before release. However, in the spring <strong>of</strong><br />

1979, most <strong>of</strong> the knapweed shoots in the treated plots were galled, although those<br />

appearing later in the summer were not infected. The plots were sampled at the end <strong>of</strong>the<br />

growing season, the knapweed counted. weighed, and the nematodes extraced (Table<br />

22). The results show that, apart from distorting the knapweed stems, the nematode had a<br />

negligible effect. This is in contrast to the results reported from the USSR (Kovalev et 01.<br />

1973). The other notable feature about the results is that in spite <strong>of</strong> the close proximity <strong>of</strong><br />

the treated and control plots, there had been little movement <strong>of</strong> the nematode into the<br />

control plots.<br />

This trial was terminated at the end <strong>of</strong> 1979, as the knapweed on the surrounding area<br />

has been treated with glyphosate and apparently eliminated.<br />

Effect <strong>of</strong> P. picridis Kirj. & Ivan. on A. repetls (L.) DC. at Jenner, Alberta in 1979.<br />

No. stems! No. seed headsl Dry weight <strong>of</strong> No. nematodes!<br />

Treatment 0.25 ml 0.25 m l fohage/0.25 m l 0.25 m:<br />

P. picridis 25.3 ± 3.9 73 ± 20.7 42 ± 6.2 2259 ± 163<br />

Control 33.2 ± 4.3 55 ± 16.0 44.8 ± 7.0 5.7 ± 3.9


Evaluation <strong>of</strong> Control Attempts<br />

Recommendations<br />

Literature Cited<br />

Acropli/on repms (L.) DC.. 109<br />

Fears were expressed following submission <strong>of</strong> the laboratory screening tests that if P.<br />

picridis was able to adapt to attack desirable knapweeds or even globe artichoke. it<br />

might be extremely difficult to eradicate. The results from the persistence trials and the<br />

lack <strong>of</strong> movement <strong>of</strong> the nematode from the treated to control plots indicate that these<br />

fears are groundless. Removal <strong>of</strong> the host plant for two growng seasons will result in<br />

eradication <strong>of</strong> the nematode with little likelihood <strong>of</strong> reinfestation from surrounding<br />

populations without human assistance. Furthermore, the host range under field conditions<br />

was narrower than in the laboratory. Thus, there were no galls found on Cynara<br />

scolymus in the plots in either Saskatchewan or Quebec, and at Regina stray plants <strong>of</strong><br />

Carduus nulans and Cirsium arvense that appeared in the treated plots were also not<br />

infected. Lettuce was not a host under laboratory conditions (Watson 1975) and the two<br />

small swellings were almost certainly from some other cause, as no nematodes were<br />

found in them. From a safety point <strong>of</strong> view. the nematode should be no problem to any<br />

plant except Russian knapweed and it can be easily eradicated if desired.<br />

The effectiveness <strong>of</strong> P. picridis is another matter. The lack <strong>of</strong> movement in or on the<br />

soil suggests that it would have to be spread thoroughly over a knapweed stand in much<br />

the same manner as a granular herbicide. The nematode also requires moist spring<br />

conditions and there is no purpose in using it on the Canadian prairies unless the area to<br />

be treated normally receives a good winter snow cover. The nematode requires a winter<br />

before it will infect Russian knapweed, so the fall seems to be the best time for<br />

application.<br />

The lack <strong>of</strong> knapweed suppression in the plots treated in Alberta was disappointing.<br />

Almost certainly the effect would have been larger if the knapweed had some grass<br />

competition, as this would have helped suppress the later growing shoots. However, the<br />

main reason for lack <strong>of</strong> an effect is that the impact may be delayed for a year. The stem<br />

growth in the current year depends on the root reserves <strong>of</strong> the previous year. The<br />

nematode creates a metabolic sink in the stem that will tend to divert reserves from the<br />

roots so the plant should be less robust for the following year.<br />

1 The nematode, P. picridis, should be released on selected major natural infestations<br />

<strong>of</strong> Russian knapweed in British Columbia, Saskatchewan, and Manitoba.<br />

2 Since Russian knapweed populations in <strong>Canada</strong> do not reproduce extensively<br />

from seed, the importation <strong>of</strong> biological control agents that attack the seed head is not<br />

justified.<br />

3 Since biological weed control programmes usually require more than one biological<br />

control agent to be successful (Harris 1979) and since P. picridis may not provide<br />

sufficient stress to control Russian knapweed in all habitats, studies could be initiated to<br />

evaluate the potential <strong>of</strong> the stem gall former Aulacida acroplilonica and the potential <strong>of</strong><br />

the rust Puccinia acroplili as possible additional biological control agents <strong>of</strong> Russian<br />

knapweed.<br />

Agriculture <strong>Canada</strong> (1967) Seeds Act and Regulations. Ottawa: Quecn's Printer. 50 pp.<br />

Alley, H.P.; Humburg. N.E. (1979) Research in weed science. 1978. AgricullUral Experiment Station <strong>of</strong> the University <strong>of</strong> Wyoming<br />

Research Journal 137.98 pp.<br />

Frazier. J.C. (1944) Nature and rate <strong>of</strong> development <strong>of</strong> the root system <strong>of</strong> Centaurea picris. Botanical Gazelle (Chicago) 105.345-351.<br />

Groh. H. (1940) Turkestan alfalfa as a medium <strong>of</strong> weed introduction. Scienct' in Agriculture 21. 36-43.


Blank Page<br />

111


Blank Page<br />

112


Pest Status<br />

Background<br />

Chapter 29<br />

Artemisia absinthium L., Absinth<br />

(Compositae)<br />

M.G. MAW and D. SCHROEDER<br />

Artemisia absinthium L. is indigenous to temperate Europe and Asia and was introduced to<br />

North America as a medicinal and flavoring herb some time before 1832. By 1841.<br />

absinth had escaped from gardens and was recognized as an established weed in the<br />

United States (Mitich 1975). The weed was first recorded in <strong>Canada</strong> at Fort Garry.<br />

Manitoba in 1860 (Scoggan 1957). By 1883, absinth was considered to be naturalized in<br />

numerous locations from Newfoundland to western Ontario (Mitich 1975). Although it is<br />

abundant in the Prairie Provinces. it was not recognized as a serious weed until 1954<br />

(Frankton & Mulligan 1970).<br />

Absinth can now be found in all provinces <strong>of</strong> <strong>Canada</strong> and in many <strong>of</strong> the northern<br />

states but with few exceptions it is confined to farm yards, around grain elevators, and<br />

along the edges <strong>of</strong> fields, railways, and roadsides. It is controlled by cultivation but can<br />

be spread into pastures, hayfields, and crop lands. Rapid increase in the spread <strong>of</strong> the<br />

weed on the prairies is associated with relatively moist conditions (Selleck & Coupland<br />

1961).<br />

Although absinth may flavor milk, and grain from infested fields may be rejected<br />

because the flour might be tainted (Selleck & Coupland 1961), such occurrences are<br />

rare. Cattle will not graze on absinth by choice, but may inadvertently consume it in hay.<br />

Absinth is a health hazard as the odor can cause illness in those working in infested<br />

fields. and its pollen causes great discomfort in those sensitized by it.<br />

While absinth is controlled by cultivation, the farming techniques <strong>of</strong> minimal or zero<br />

tillage might permit the weed to become a serious problem in some areas.<br />

At <strong>Canada</strong>'s request. surveys <strong>of</strong> absinth insects in Europe were undertaken by the<br />

Commonwealth Institute <strong>of</strong> Biological Control. A literature review suggested that<br />

Euzophera citlerosella (Zeller) (Lepidoptem: Pyralidae), a stem and root miner, was<br />

specific to A. absimhillm (Miotk 1973). Field surveys in Austria, Germany, France, and<br />

the Swiss Valais showed that E. cinerosel/a occurs in most populations <strong>of</strong> absinth<br />

throughout its European mnge, and that it causes serious damage to its host (Schroeder<br />

1979).<br />

E. cillerosella emerges about the end <strong>of</strong> May to the end <strong>of</strong> July and mating takes place<br />

within 24 hours <strong>of</strong> the female's emergence. Males mate several times; females just once.<br />

Oviposition starts the day following mating with 12 eggs laid daily for the first four days.<br />

Over 100 eggs may be laid during June and half that number during July.<br />

The eggs arc deposited on the lower leaves and hatching occurs in 8-10 days. The<br />

young larvae commence to feed on the upper cuticle <strong>of</strong> the leaf but soon bore into the<br />

leaf petiole destroying the bud in the leaf axil, then downwards, feeding on the vascular<br />

tissue <strong>of</strong> the shoots. As the season advances, the larvae reach the roots and destroy<br />

the outer parts before mining into the woody tissue. The winter is passed as hibernating<br />

larvae and puplltion takes pl.lce in the spring within a silken cocoon in a pupal cell near<br />

the root crown.<br />

Damllge to the host plant depends on the development <strong>of</strong> the larvae. First and second<br />

instars clluse little damage. but third and fourth instars mine the cambium and vascular<br />

113


Pest Status<br />

Chapter 30<br />

Carduus nutans L., Nodding Thistle and<br />

C. acanthoides L., Plume less Thistle<br />

(Compositae)<br />

P. HARRIS<br />

Carduus nwaftS L. and C. acanlllOides L. are thistles <strong>of</strong> European origin that form dense<br />

stands on dry uncultivated grasslands in many parts <strong>of</strong> North America. Three interbreeding<br />

subspecies <strong>of</strong> C. nWans occur in <strong>Canada</strong> (Moore & Frankton 1974) which arc<br />

treated as full species by some authors (Kazmi 1964). There is also some hybridization<br />

between C. mllaftS and C. aca""lOides (Mulligan & Frankton 1954). however. as far as<br />

the practical aspects <strong>of</strong> biological control are concerned there are two species: C. nulaftS<br />

with large heads (2-7 cm diameter) and a relatively short but intense flowering period<br />

that, on Saskatchewan rangeland, is largely finished by the end <strong>of</strong> July; C. aca",hoides<br />

with small heads (1.2-2.5 cm diameter) and a prolonged flowering period that. in<br />

southern Ontario. produces a succession <strong>of</strong> flower heads from early June until the plant<br />

is killed by frost in October or later. The seed <strong>of</strong> both thistles germinates as conditions<br />

permit throughout the growing season and they may be summer annuals. winter annuals. or<br />

biennials. The flowering height in C. nWans ranges from a few centimeters to over 2 m.<br />

so there is considerable plasticity in the species as well as genetic variability as the result<br />

<strong>of</strong> hybridization.<br />

Distribution <strong>of</strong> the thistle on the Canadian prairies is limited by a need for winter<br />

protection <strong>of</strong> the rosettes and relatively low grass competition for seedling establishment.<br />

Thus it is mainly found on light soils in the mid-grass prairie vegetation zone. in<br />

places that are covered by snow drifts in winter. such as gullies. fence lines. brush<br />

patches. and the lee side <strong>of</strong> stone piles. The dead stems also trap the snow in winter.<br />

Established stands tend to be self-perpetuating as the death <strong>of</strong> the flowering stems in<br />

August creates a seedbed largely devoid <strong>of</strong> competing vegetation. The occurrence <strong>of</strong> the<br />

thistle on heavier soils is a sign that the site was disturbed within the past few years.<br />

The status <strong>of</strong> C. nwaftS has changed in Saskatchewan since the establishment <strong>of</strong> the<br />

seed-head weevil. Rhinocyllus conicllS Froel.. for its biological control. In 1970, stands<br />

<strong>of</strong> 150 OOOlha <strong>of</strong> flowering C. nWans plants were common in pasture and extended for<br />

many kilometers in roadside borrow pits. At this density. there is no grazing within the<br />

stand. In 1980. the thistle was largely restricted to breaks in the pasture sward such as<br />

ground squirrel diggings and along cattle trails. Stands <strong>of</strong> over 15 000 plantslha were<br />

uncommon and small roadside stands that used to be a feature <strong>of</strong> the landscape have<br />

been reduced to scattered plants. On the 4 point scale. modified from Vere & Medd<br />

(1979). the most densely infested region has declined from category 4 to categories 1-2:<br />

1 = no C. nmans, 2 = scattered thistles or < I stand/ha. 3 = scattered thistles with 1 to 50<br />

standslha.4 = continuous thistles or >50 stands/ha. There was negligible grazing loss<br />

from category 2 infestations in Australia. a loss <strong>of</strong> 8.3% from category 3. and 16.7% loss<br />

from category 4. No comparable data are available for Saskatchewan but the loss is<br />

presumably similar. Currently there is little grazing loss from C. mllaftS on permanent<br />

pasture although dense patches remain on recently disturbed and abandoned ground<br />

which is <strong>of</strong>ten on roadsides. At low densities it is still a nuisance in some places such as<br />

parks. C. nmans continues to spread in the mid-grass prairie zone but on a narrower<br />

range <strong>of</strong> sites.<br />

115


Releases and Recoveries<br />

Rhinocynus conicus<br />

(Froel.) (Coleoptera:<br />

Curculionidae) (a) Ecology<br />

Trichosirocalus<br />

horridus (panz.)<br />

(Coleoptera:<br />

Curculionidae)<br />

Carduus IIutans L., 117<br />

The biology and host specificity <strong>of</strong> R. conicus has been treated in detail by Zw61fer &<br />

Harris (in press). The weevil is native to Europe, western Asia, and North Africa<br />

between latitudes 30 0<br />

N and SooN. Its main host is C. nutans but there has been race<br />

formation on other thistles in the genera Carduus, Cirsium, Silybum, and Onopordum.<br />

Teneral R. conicus emerge in July to early August in southern Ontario and August in<br />

Saskatchewan. The weevil has a second generation if the day length is over 16 hrs.<br />

Individuals released in a cage in southern Ontario on the 27 June when the day length was<br />

15 hr 48 min did not breed, but those released in mid-July in Saskatchewan with a day <strong>of</strong><br />

16 hr 5 min did so. Laing & Heels (1978) reported a second generation from weevils<br />

released in southern Ontario in August, but these were mated in the laboratory. presumably<br />

under long day conditions. Normally, in <strong>Canada</strong>, there is a single generation emerging<br />

after mid-summer and then hibernating in the soil litter. They appear again in the spring<br />

to feed on the leaves <strong>of</strong> bolting thistles and oviposit on the involucral bracts <strong>of</strong> the flower<br />

buds. Oviposition starts slightly before the buds are available in the spring with some<br />

wastage <strong>of</strong> eggs on the leaves enclosing the terminal flower bud and the first heads receive<br />

an abundance <strong>of</strong> eggs. Rees (1977) counted over 500 eggs on a single head.<br />

The eggs are covered with a cap <strong>of</strong> chewed thistle and the larva bores directly into the<br />

flower bud from under the cap. It mines the receptacle and sometimes the peduncle as<br />

well as feeding on the young ovules. Typically the receptacle mine fills with callus on<br />

which the larva feeds (Shorthouse, 1982, personal communication). Thus, like a gallformer,<br />

the larva stimulates production <strong>of</strong> its food supply. The mature larva forms a<br />

hard pupal chamber in the thistle head which remains after emergence so the number<br />

that developed in a head can be counted.<br />

(b) Releases<br />

Most <strong>of</strong> the R. conicus stock established in <strong>Canada</strong> was collected from C. nwans growing in<br />

the French Rhine Valley around Mulhouse. As the initial results with this stock on C.<br />

acanthoides were disappointing, additional weevils were imported from the smallheaded<br />

thistle C. personata (L.) Jacquin as well as stocks from Cirsium spp. Day length<br />

at the release date probably prevented many <strong>of</strong> the colonies from breeding until the year<br />

following release (Table 23). With hindsight, it would have been better to have held them<br />

in laboratory storage until the following spring or mated them under long day conditions<br />

in the laboratory before release.<br />

(a) Ecology<br />

The biology <strong>of</strong> T. horridus was reported by Trumble & Kok (1979) and a bibliography<br />

<strong>of</strong> the weevil compiled by Trumble & Kok (1980). The weevil is found from Italy to<br />

Poland, in Turkey and the USSR. The adults have been found feeding on a number <strong>of</strong><br />

thistle genera and the larvae developed on several <strong>of</strong> these in the laboratory (Kok 1975),<br />

however, in the thistle garden at the Regina Research Station, they have only been found<br />

on Carduus spp. and not on either European or native Cirsium spp. Thus the field host


118 P. Harris<br />

Table 23 Open releases and recoveries <strong>of</strong> R. conicllS Froel. against Cardlllls spp.<br />

Release Initial<br />

Province Host date Place Source Number recovery<br />

Saskatchewan C. "Ulans 26.7.68 Aylesbury 400 1969<br />

16.7.69 1898<br />

C. "Ulans 16.7.69 Findlater 1897 1969<br />

Ontario C. acamhoides 25.7.68 Read (site A) 320 1969<br />

23.5.69 1 282<br />

10.6.70 2 350<br />

5.10.69 Read (Site B) 3 74<br />

5.10.69 1 80<br />

31.6.70 Read (Site C) 1 3500 1971<br />

22.5.69 Stoco Lake 1 350 1970<br />

11.7.69 Roslin I 4320 1970<br />

22.5.70 1 1500<br />

1.5.70 Moira 4 200<br />

C. acamhoides 3.6.70 Flesherton 1 250 not checked<br />

xc. "utans<br />

C. "ulans 9.6.75 Guelph 5 2185 1976<br />

Manitoba C. "Ulans 17.6.74 Somerset 5 500 1975<br />

4.9.74 Darlingford 5 4000 1975<br />

Quebec C. "ulans 17.6.74 Lac St-Jean 5 1100 1975<br />

C. acanthoides 23.7.78 Huntingdon 5 7731 1982<br />

16.6.80 5 10130<br />

British Columbia C. nUlans 16.7.79 Williams Lake 5 2600 site<br />

destroyed<br />

1. Mulhouse, France ex C. nutans<br />

2. St. Hippolyte, France ex C. personala<br />

3. Krymsk, USSR ex Cirsium spinosllm<br />

4. Nantes, France ex C. vulgare<br />

5. Findlater. Saskatchewan ex C. nutans<br />

range <strong>of</strong> the strain introduced is more restricted than that <strong>of</strong> R. conicus. Under field<br />

conditions in Virginia, United States, it had a strong preference for C. nutans over C.<br />

acanlhoides and was more readily established on stands <strong>of</strong> the former species (Sieburth<br />

& Kok 1982). On C. nUlans, high densities <strong>of</strong> T. IIOTTidus and R. conicus were found in<br />

the same stands (Kok 1980).<br />

In <strong>Canada</strong>, the adult weevil overwinters in the soil litter. It becomes active in early<br />

spring to lay in the mid-vein <strong>of</strong> large CardullS leaves. This differs from the life cycle in<br />

Rome, Italy, where the weevil starts ovipositing in October and continues over winter<br />

(Bolt & Campobasso 1981). The larva bores down the mid-vein into the crown and<br />

develops just below the apical leaves. This causes the plant to produce several lateral<br />

shoots, which in Saskatchewan may be attacked again when about 15 cm high; the larvae<br />

are also sometimes found in the lateral vegetative buds. The result is to produce a bushy<br />

thistle in which flowering is probably delayed. The larvae pupate in the soil and emerge<br />

as adults in reproductive diapause in the early summer. They feed on the thistle leaves<br />

until the fall and can be frequently found on the bracts <strong>of</strong> the flower buds.<br />

(b) Releases<br />

The stock released at Aylesbury, Saskatchewan, was collected in eastern Austria. around<br />

Morel, Switzerland, and Neuenburg, Germany (Table 24). In addition, 22 weevils from


Table 24<br />

Evaluation <strong>of</strong> Control Attempts<br />

Carduus IIll1allS L., 119<br />

Neuenburg, released on a 4 m 2 stand <strong>of</strong> C. IIll1allS maintained at the Regina Research<br />

Station, were the source <strong>of</strong> the stock released elsewhere in <strong>Canada</strong>. At Regina, survival<br />

<strong>of</strong> the weevil was most easily confirmed by removing the apical rosette leaves to expose<br />

the larvae during the spring wheat seeding period, usually around mid-May. Later in the<br />

year, the weevil was hard to find and even its damage was obscure. Thus, it was hard to<br />

confirm establishment at Aylesbury as the road to the site was <strong>of</strong>ten impassable until late<br />

May. It is suspected that the reason that this cryptic weevil has not been reported<br />

established at other sites is that the thistles have been inspected too late in the season.<br />

Open releases and recoveries <strong>of</strong> T. horridlls (Panz.) against CardulLf spp.<br />

Province Site Host Year Source Number Recovery<br />

Saskatchewan Aylesbury C. flUlaflS 1975 Austria-Germany-<br />

Switzerland<br />

65 1979<br />

Aylesbury C. flUians 1979 Saskatchewan 36<br />

Regina C. nUians 1975 Germany 22 1976<br />

Quebec Huntingdon C. acanthoides 1977 Saskatchewan 26<br />

Huntingdon C. acallthoides 1980 Saskatchewan 6<br />

Ontario West<br />

Huntingdon C. acallthoides 1978 Saskatchewan 33<br />

British<br />

Columbia<br />

Williams<br />

Lake C. nutans 1979 Saskatchewan 22<br />

site<br />

destroyed<br />

Manitoba Snowflake C. nutans 1980 Saskatchewan 16<br />

Density <strong>of</strong> R. conicus<br />

The general effects <strong>of</strong> R. conicus on C. lIlllans in Saskatchewan have been described in<br />

the Pest Status section. More detailed studies were done at the Findlater and Aylesbury<br />

release sites (Table 25 and 26). Findlater was an unused gravel pit and Aylesbury a gully in<br />

a permanent pasture. Both sites were sampled in early August: Findlater on a permanent<br />

transect along the spoil pile, and Aylesbury in four directions from the release point. The<br />

density <strong>of</strong> flowering thistles was determined with m 2 samples taken at 5 or 10 m intervals,<br />

or they were calculated by the distance between nearest neighbour for a population with a<br />

uniform distribution (Southwood 1978). The latter estimates are indicated by the absence<br />

<strong>of</strong> an error term in Table 26. In 1978 and 1979, the estimates obtained from the nearest<br />

neighbour method were: Aylesbury 1.4, 1.0, and at Findlater 48.2, 8.3/m 2 • These are<br />

similar to the values obtained from counting on the m 2 plots (Table 26). However, in 1976<br />

at Aylesbury, the distribution was clumped with most <strong>of</strong> the site free <strong>of</strong> thistles but some<br />

patches containing up to 651m 2 • In this year the nearest neighbour method, assuming a<br />

uniform distribution, gave a density <strong>of</strong> 2.7/m 2 compared to 6.6/m 2 obtained by averaging<br />

the m l samples. All heads were clipped from the thistle nearest to the right comer <strong>of</strong> the<br />

plot, the receptacle diameter <strong>of</strong> each head was measured, and the number <strong>of</strong> R. conicus or<br />

its pupal chambers were counted. Ifthere were not enough thistles in the transect samples<br />

to determine weevil density, the thistle nearest to each sampling interval was used.<br />

At Findlater, the density <strong>of</strong> R. coniClLf maturing in the flower heads climbed steadily<br />

from the second year to reach a plateau <strong>of</strong> 3.4 to 6.4 weevilslhead (Table 25). The weevils<br />

released in 1968 at Aylesbury (Table 23) received a day <strong>of</strong> less than 16 hrs in length and<br />

so probably did not breed until the following year. Thus the initial increase at Aylesbury<br />

was similar to that at Findlater; but after reaching a peak in 1975, the density fell. This


120 P. Harris<br />

Table 25<br />

may have been caused by cattle consuming the terminal flower heads, which contained<br />

the highest numbers <strong>of</strong> R. conicus. The cattle did this to an increasing extent as the<br />

density <strong>of</strong> the stand declined. The site was not grazed in 1982 and the density <strong>of</strong> weevils<br />

increased.<br />

Establishment pattern <strong>of</strong> R. conicus Froel. on C. acantlloides L. and C. nutans L.<br />

C. acantlloides (Site A) at Read, Ontario. Released 320 R. conicus 25 July 1968,282 on<br />

23 May 1969, 350 on 10 June 1970.<br />

Calc. R. coniclls<br />

Year Eggsffhistle per head No. thistles sampled<br />

1968 0 0<br />

1969 0.29 0.017 1453<br />

1971 0.38 0.023 285<br />

C. acantlloides at Roslin, Ontario. Released 4320 R. conicus on 11 July 1969.<br />

Calc. R. conicus<br />

Year Eggs/thistle per head No. thistles sampled<br />

1969 0 0<br />

1970 0.52 0.03 1531<br />

1970· 1.56 0.09 50<br />

1971 0.18 0.01 1272<br />

• On scattered C. nutans in C. acantlloides stand.<br />

C. nutans at Aylesbury, Saskatchewan. Released 400 R. conicus 26 July 1968, 1898 On<br />

16 July 1969.<br />

Year R. conicuslhead No. heads sampled<br />

1969 0.08 317<br />

1970 0.04 1055<br />

1971 0.06 688<br />

1972 0.26 847<br />

1973 0.79 956<br />

1974 2.72 311<br />

1975 3.68 288<br />

1976 1.43 182<br />

1977 0.43 353<br />

1978 2.70 251<br />

1979 0.86 349<br />

1980 0.40 293<br />

1981 0.06 54<br />

1982 3.18 105


Table 25<br />

continued<br />

CtmillllS IIlllatl.f L.. 121<br />

C. ,wlans at Findlater. Saskatchewan. Released 1897 R. conicus on 16 July 1969.<br />

Year R. conicllslhead No. heads sampled<br />

1969 0.03 119<br />

1970 0.002 509<br />

1971 0.01 194<br />

1972 0.19 241<br />

1973 0.46 168<br />

1974 1.45 211<br />

1975 4.53 52<br />

1976 4.01 96<br />

1977 1.84 99<br />

1978 3.38 86<br />

1979 5.02 177<br />

1980 3.65 373<br />

1981 3.51 59<br />

1982 6.36 375<br />

Laing & Heels (1978) reported achieving similar population densitites on C. n!llans<br />

in Ontario and the initial results reported by Letendre el al. (1976) for Quebec followed<br />

the same pattern. Thus the Saskatchewan impact <strong>of</strong> R. coniclls on C. nWans is probably<br />

representative <strong>of</strong> populations elsewhere in <strong>Canada</strong> except for British Columbia where the<br />

release site was destroyed by road building.<br />

Density <strong>of</strong> C. tllllans<br />

The density <strong>of</strong> C. tlllIans declined at Aylesbury from 15.7 flowering plantslm 1 in 1969 to<br />

0.5/ml in 1982 (Table 26). The habitat was initially invaded by Russian pigweed, Axyris<br />

amarantlroides L., and stinkweed, Thlaspi arvense L. The stinkweed was cropped by<br />

the red turnip beetle. Entomoscelis americana Brown, which also developed on C.<br />

mllans rosette leaves during this transitional period. Finally perennial grasses. Agropyron<br />

repens (L.) Beauv. and Sripa comata Trin. & Rupr.. occupied the area and the<br />

thistle was confined to breaks in the sward.<br />

At Findlater. the density <strong>of</strong> the thistle fluctuated from dense. following a moist fall<br />

and/or spring. to sparse when it was dry (Table 26). The invasion <strong>of</strong> the transect on the<br />

spoil pile by other plants was slow, so after dry winters the site remained open for<br />

recolonization by C. nlllans when conditions were more favourable for its seedlings.<br />

Gradually clumps <strong>of</strong> crested wheat grass, Agropyron crisralllm (L.) Gaertn .• appeared<br />

and by 1982 covered the transect. The C. nWans stand was then confined to the southwestern<br />

slope <strong>of</strong> the spoil pile. which was still free <strong>of</strong> grass.<br />

The decline <strong>of</strong> C.nlllans stands in Virginia, United States. occurred more rapidly than<br />

in Saskatchewan (Kok & Surles 1975). The summer rainfall was greater and its distribution<br />

more even so that there would be a greater growth <strong>of</strong> grass. Thus both studies<br />

support the conclusion that control <strong>of</strong> C. nutans has been achieved by a combination <strong>of</strong><br />

R. conicus and competition from other plants.<br />

At neither Saskatchewan site has the lowered density <strong>of</strong> the thistle reduced the<br />

numbers <strong>of</strong> R. cotliells/head. Scattered thistles <strong>of</strong>ten escape attack and the weevil may<br />

not be able to maintain itself on them. For example, 120 R. conieu.s released in 1974 on a<br />

stand <strong>of</strong> C. mllans in Regina that consisted <strong>of</strong> 368 scattered phtnts at 0.07 plantslm 1 bred<br />

in 1975 but not in 1976 when the density was 0.02 plantslm 1 • However, the weevil


122 P. Harris<br />

Table 26 C. flU/ailS L. stands at Aylesbury and Findlater following establishment <strong>of</strong> R. coniclls<br />

Froel. (± S.E.M.).<br />

Year No. C. tIIltallS/m 2 No. heads/plant Head diameter No. plants sampled<br />

Aylesbury<br />

1969 15.9 ± 1.7 5.7 ± 0.8 2.87 ± 0.02 56<br />

1970 11.8 ± 1.1 7.5 ± 0.9 1.68 ± 0.03 125<br />

1971 6.8 ± 0.7 6.2 ± 0.8 2.02 ± 0.03 100<br />

1972 604 ± 0.8 8.0 ± 0.9 1.65 ± 0.02 106<br />

1973 704 ± 1.1 1.79 ± 0.02 129<br />

1974 4.1 6.6 ± 1.3 1.73 ± 0.02 47<br />

1975 4.9 404 ± 0.7 1.94 ± 0.02 65<br />

1976 6.6 ± 104 304 1.73 ± 0.01 54<br />

1977 1.3 ± 0.3 6.0 ± 0.3 1.55 ± 0.02 60<br />

1978 1.2 ± 0.5 4.3 ± 0.8 1.52 ± 0.03 58<br />

1979 104 ± 004 3.9 ± 0.7 1.38 ± 0.02 89<br />

1980 1.9 ± 0.7 2.3 ± 0.1 1.16 ± 0.05 124<br />

1981 0.9 ± 0.4 1.5 ± 0.1 1041 ± 0.02 60<br />

1982 0.5 ± 0.1 2.8 ± 0.5 1.56 ± 0.04 38<br />

Findlater<br />

1969 l3.5 ± 2.6 11.9 ± 1.8 1.80 ± 0.04 10<br />

1970 19.1 ± 2.5 10.0 ± 1.3 1.98 ± 0.02 51<br />

1971 19.8 ± 2.8 4.9 = 0.5 1.69 ± 0.04 40<br />

1972 8.0 ± 1.2 6.9 = 0.7 1.70 ± 0.02 35<br />

1973 5.3 = 1.0 1.87 ± 0.04 32<br />

1974 19.9 5.9 ± 1.2 1.95 ± 0.01 35<br />

1975 52.1 2.9 ± 0.7 1.86 ± 0.06 18<br />

1976 7.2 4.2 ± 0.6 1.81 ± 0.03 24<br />

1977 0.4 ± 0.2 5.2 ± 0.2 2.03 ± 0.03 19<br />

1978 33.9 ± 5.7 2.3 ± 0.3 lAO ± 0.05 38<br />

1979 6.7 ± 0.1 5.2 ± 0.8 1.19 ± 0.03 33<br />

1980 3.0 ± 0.7 3.6 ± 0.3 1.19 ± 0.03 104<br />

1981 0.6 ± 0.3 4.0 ± 1.6 1.39 ± 0.06 24<br />

1982 0.03- 11.6 ± 1.8 1.83 ± 0.02 32<br />

- 1 thistle in 37 m 2<br />

persisted at another site in Regina on a 10 m 2 stand <strong>of</strong> C. nlltallS with around SO flowering<br />

thistles. If the weevil requires dense stands, even if they are confined to a few m 2 , the<br />

roadside and gravel pit stands may be necessary for maintaining a population.<br />

Size <strong>of</strong> C. nU/alls<br />

The size <strong>of</strong> flowering C. 'IlltallS plants as indicated by the number <strong>of</strong> heads/plant<br />

declined (Table 26). The slope <strong>of</strong> the decline was not significantly different at Aylesbury<br />

and Findlater, so the data were combined for the regression in Table 27 (Equation 1).<br />

The only major departure from the regression was the 1982 value from Findlater. This<br />

value was not used for the regression as it was not from the permanent transect. The<br />

equation shows that on the average a plant produced 0.34 fewer heads each year. Thus


Table 27<br />

Table 28<br />

Carduus lIU1tl1lS L. . 123<br />

thistle size has declined more than its density. For example at Aylesbury. the density <strong>of</strong><br />

the thistle in 1982 was 97% less than in 1969. but the number <strong>of</strong> heads had declined by<br />

nearly 99%.<br />

Regression equations <strong>of</strong> the effect <strong>of</strong> R. con;cus Froel. on C. nUiallS L.<br />

1. Average no. heads/plant at Aylesbury<br />

and Findlater<br />

2. Average head diameter at Aylesbury<br />

3. Average head diameter at Findlater<br />

x = no. years r = correlation coefficient<br />

dJ.<br />

= 994.5 - 0.50 x 26<br />

= 79.8 - 0.40 x 12<br />

7.3 - 2.89 x 13<br />

r<br />

-0.79<br />

-0.71<br />

-0.38<br />

p<br />


124 P. Harris<br />

Reasons for the impact <strong>of</strong> R. conicus<br />

Lashley (1969) and Sagar (I972) calculated that over 98% <strong>of</strong> the seed would have to be<br />

destroyed to achieve control <strong>of</strong> C. nutans if 50% <strong>of</strong> the seed was viable and 10% <strong>of</strong> the<br />

plants survived. At Aylesbury 80.6% ± 3.6 <strong>of</strong> the plump seed germinated on moist filter<br />

paper at room temperature, so the viability factor they used was low. Regression<br />

equations <strong>of</strong> seed production per head versus head size and the number <strong>of</strong> R. conicus<br />

gave reductions that varied from 5 to 25 seeds per weevil. The variation may arise from<br />

the formation <strong>of</strong> callus tissue acting as a metabolic sink in the heavily attacked heads;<br />

but even the highest estimate would only reduce seed production by around 50%.<br />

Analysis <strong>of</strong> one year's results <strong>of</strong> a four year study <strong>of</strong> thistle survival at Aylesbury<br />

indicated that no seedlings survived in a grass sward. Presumably the stand became<br />

established initially after a natural or man-made disturbance exposed the ground. The<br />

thistle then perpetuated itself and even at the stand margin spread by producing enough<br />

plants to smother the site. A constant percentage <strong>of</strong> plants was lost through the growing<br />

season regardless <strong>of</strong> their density or age and around 3% survived to the flowering stage.<br />

Apparently the weevil reduced seed production to the point at which the thistle was not<br />

able to completely occupy the site at all times. This allowed the entry <strong>of</strong> competing<br />

vegetation which eventually displaced the thistle. Further reductions <strong>of</strong> C. nUians can<br />

almost certainly be achieved by pasture improvement to reduce the number <strong>of</strong> bare spots.<br />

Effect <strong>of</strong> R. coniclls on C. acanthoides<br />

C. acanthoides has remained a problem in Ontario since the introduction <strong>of</strong> R. conicus. It<br />

is known that the weevil has persisted on the thistle as there is no difficulty in finding<br />

flower heads with eggs on them in the early spring; however, it is probable that seed<br />

production is reduced by a minor amount. In Virginia, only 12.4% <strong>of</strong> C. acanthoides<br />

heads are attacked compared to 81.4% <strong>of</strong>those on C. nlltans (Surles & Kok 1977). Much<br />

<strong>of</strong> this difference relates to the synchronization <strong>of</strong> egg production with flowering. In<br />

Saskatchewan, oviposition covers the flowering period and both are largely completed in<br />

July. In contrast, on 7 August 1969 at Read, Ontario, after R. conicus had ceased<br />

oviposition, 22% <strong>of</strong> the heads <strong>of</strong> C. acanthoides had finished flowering and would have<br />

been available to R. conicus, 7% were in bloom, and 70.7% were still in bud (unpublished<br />

data, T. New). Thus, 78% <strong>of</strong> the heads produced by C. acanthoides were not available to<br />

R. conicus. Similarly in Virginia, Surles & Kok (1977) found 19.1% <strong>of</strong> the heads were<br />

available to the weevil.<br />

Dowd & Kok (1981) found a density-related weight reduction <strong>of</strong> R. conicus in C.<br />

acanthoides that did not occur in C. n14tans. Also, the weevil responded to C. nutans<br />

more readily than to C. acanthoides for oviposition: in Table 25 there was a 3-fold<br />

difference in the numbers <strong>of</strong> eggs laid on the two species while Surles & Kok (1977)<br />

found a 4-fold difference. The difference in flower head size does not seem to be the<br />

important factor as some races <strong>of</strong> R. conicus achieve high densities on the small-headed<br />

thistles, C. pycnocephalus and C. ten14iflorus.<br />

R. conicus has not adapted to C. acanthoides in Europe: Batra et al. (1981) found it<br />

rarely on the thistle in France and it was not recorded in the surveys on C. acanthoides<br />

by Zw6lfer (1965). In my opinion, it is unlikely to adapt to become an effective control<br />

agent <strong>of</strong> the thistle in North America.<br />

Effect <strong>of</strong> T. horrid/IS<br />

No assessment <strong>of</strong> the impact <strong>of</strong> T. horridus on C. nil tans was made at Aylesbury beyond<br />

the fact that few plants are attacked. There are no reports on the value <strong>of</strong> the more dense


126 P. Harris<br />

Laing. J.E.: Heels. P.R. (1978) Establishment <strong>of</strong> an introduced weevil. Rhinoc}'lIlls conicllS (Coleoptera: Curculionidae) for the biological<br />

control <strong>of</strong> nodding thistle. Cardull.f nlltans (Composit3e) in southern Ontario. Proceedings <strong>of</strong> the<br />

ElIlomological Societ}' <strong>of</strong> On",rio 1119. 3-8.<br />

Lashley. R. (1969) Musk thistle control. Proceedings <strong>of</strong> Ihe North Ctlltral Weed Control Conference 24. 99-100.<br />

Letendre. M.e.: Ritchot. 1'01.: Guibord. O·e.: Leduc. e. (1976) Essai d'cradication du chardon penche. CardullS nutans L.. oj I'aide du<br />

charanllon Rhinoc}'lIlls conicus Fmc!. Ph}'toprotection 57. 47-54.<br />

Moore. R.J.: Frankton. C. (1974) The thistles <strong>of</strong> <strong>Canada</strong>. Canadian Department <strong>of</strong> Agriculture Monograph 10. Ottawa. III pp.<br />

Mulligan. G.A.: Frankton. e. (195-1) The plumeless thistles (Cardlms spp.) in <strong>Canada</strong>. Canadian Field·Naluralist68. 31-36.<br />

Myers. J.H.: Harris. P. (1980) Distribution <strong>of</strong> Urophora galls in nower heads <strong>of</strong> diffuse and spotted knapweed in British Columbia. Jourtull<strong>of</strong><br />

Applied Ecolog}' 17.359-367.<br />

Puttler. B.: Long. A.H.: Peters. E.J. (1978) Establishment in Missouri <strong>of</strong> Rhinocyllus conicllS for the biological control <strong>of</strong> musk thistle<br />

(Carduus nu",IIS). Wud Science 26. 188-190.<br />

Rees. N.E. (l9n) Impact <strong>of</strong> RhinoC)'lIus COIliCIlS on thistles in southwestern Montana. Em'ironmental Entomology 6. 839-842.<br />

Rees. N.E. (1978) Interactions <strong>of</strong> Rhinocyllus ccmicus and thistles in the Gallatin valley. 31·38. In: Frick. K.E. (Ed.) Biological control <strong>of</strong><br />

thistles in the genus Card,ms ill the United States. Stoneville. Miss.: USDA. 50 pp.<br />

Sagar. G.R. (1972) On the ecology <strong>of</strong> weed control. In: Jones D.P.: Solomon M.E. (Eds.) Biology in pest and disease control. British<br />

Ecological <strong>Society</strong>. 42-50.<br />

Siebunh. P.J.: Kok. L.T. (1982) Oviposition preference <strong>of</strong> Tricllosirocalus Irorridus (Coleoptera: Curculionidae). Canadian Entomologist<br />

114. 1201-1202.<br />

Southwood. T.R.E. (1978) Ecological methods. Chapman & Hall. 52-1 pp.<br />

Surles. W. W.: Kok. L.T. (1977) Ovipositional preferences and s)'nchronization <strong>of</strong> Rlrinocyllus conicus with Carduus nutans and C.<br />

acantlroides. Em·ironmental ElIlomolog}, 6. 222-22-1.<br />

Trumble. J.T.: Kok. L.T. (1979) Celltorhynchidius horridus (Coleoptera: Curculionidae): life cycle and development on Carduus thistles in<br />

Virginia. Annals <strong>of</strong> tht <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> America n. 563-564.<br />

Trumble. J.T.: Kok. L.T. (1980) A bibliography <strong>of</strong> Ceuthorhynchidius horridus (Panzer) (= Trichosirocalus horridus (Panzer»). an<br />

introduced weevil for the biological control <strong>of</strong> Carduus thistles. Bulletin <strong>of</strong> the <strong>Entomological</strong> Sociery <strong>of</strong><br />

America 62.464-467.<br />

Vere. D.T.; Medd. R.W. (1979) Estimating the economic loss caused by Carduus nutans in New South Wales. Proceedings <strong>of</strong> the 7th Asian·<br />

Pacific Weed Science Conference. 421-423.<br />

Ward. R. H.; Pienkowski R.L.: Kok. L. T. (1974) Host specificit)· <strong>of</strong> the first·instar <strong>of</strong> Cellthorhynchidius horridus. a weevilforthe biological<br />

control <strong>of</strong> thistles. Journal <strong>of</strong> Economic Entomology 67, 735 -737.<br />

Zw6ICer, H. (1965) Preliminary list <strong>of</strong> phytophagous insects attacking wild Cynareae (Compositae) species in Europe. Commonwealth<br />

Instilllte <strong>of</strong> Biological Control Technical Bulletin 6. 81-154.<br />

Zw6lfer, H.; Harris, P. (in press) Biology and host specificity <strong>of</strong> Rhinocyllus conicllS, (Froel.) (Col.: Curculionidae), a successful biocontrol<br />

agent <strong>of</strong> the thistle Carduus nutans L. Zeirschrift fiir angewandte Entomologie.<br />

Zw61fer, H.; Preiss. M. (1983) Host selection and oviposition bchaviour in west· European ecotypes <strong>of</strong> Rhinocyllus conicus FroeJ. (Col.:<br />

Curculionidllc). Zeitsclrrift fiir angewandte Entomologie 95, 113·122.


Pest Status<br />

Chapter 31<br />

Centaurea diffusa Lam. and C. maculosa<br />

Lam. s. lat., Diffuse and Spotted<br />

Knapweed (Compositae)<br />

P. HARRIS and I.H. MYERS<br />

Diffuse and spotted knapweed are herbaceous plants introduced from Europe to the dry<br />

grasslands <strong>of</strong> western <strong>Canada</strong> and other parts <strong>of</strong> North America. The combination <strong>of</strong><br />

their allelopathic properties (Fletcher & Renney 1963), low forage value. and drought<br />

adaptation have allowed these two knapweeds to displace most other herbaceous plants<br />

over large areas.<br />

Diffuse knapwced is typically a biennial. The species was recorded in Washington State<br />

in 1907 (Howell 1959) and early Canadian herbarium specimens suggest a northward<br />

spread into British Columbia: Oyama 1936, Penticton 1939, Grandforks 1940 (Groh<br />

1943); however, it may also have been introduced directly into the province with Turkestan<br />

alfalfa from the Caspian Sea region. Renney (1959) suggested that the weed occurred at<br />

Pritchard and Lytton, British Columbia, prior to 1930 and certainly other weeds from<br />

southern USSR were recorded at Kamloops (near Pritchard) as early as 1920 (Groh<br />

1940). From these beginnings the weed spread by 1972 to infest 25 952 ha <strong>of</strong> dryland range<br />

and roadside in British Columbia as well as small areas in Alberta (Watson & Renney<br />

1974). Distribution <strong>of</strong> diffuse knapweed in 1977 is shown in Fig. 4. In 1981 about 25 ha <strong>of</strong><br />

diffuse knapweed, 8 heavily infested. were found on one farm near Morden, Manitoba<br />

(B. Todd, 1982, personal communication). Harris & Cranston (1979) predicted that the<br />

relatively high availability <strong>of</strong> summer moisture in Manitoba would put it beyond the range<br />

<strong>of</strong> the weed. It is now not clear whether the potential <strong>of</strong> the weed was seriously<br />

underestimated or the infestation merely represented a local site anomaly as it was on a<br />

light soil with a southern exposure and dry. In 1982 several kilometers <strong>of</strong>railway track<br />

were found to be infested at Walsh, Saskatchewan. This is only a slight extension <strong>of</strong> the<br />

range from the stand at Irvine, Alberta (Fig. 4); but a small stand was also found on the<br />

railway track at Colonsay, Saskatchewan, which is southeast <strong>of</strong> Saskatoon (S. McKell,<br />

personal communication). Distance spread <strong>of</strong>the weed across the prairies <strong>of</strong>ten seems to<br />

be associated with railway ties or treated bridge timbers from British Columbia.<br />

It appears that diffuse knapweed will continue to spread until it dominates the herbaceous<br />

community in open uncultivated sites in the brown chernozem and brunisol soils <strong>of</strong><br />

southern British Columbia and the brown soils <strong>of</strong> Alberta and Saskatchewan. This<br />

involves about 7.5 million ha (Harris & Cranston 1979).<br />

Spotted knapweed is a short lived perennial. The first Canadian specimen was collected<br />

by McCoun from Victoria, British Columbia, in 1893 (Groh 1943). By 1972, it was<br />

estimated to have infested 2410 ha in British Columbia as well as several small stands in<br />

Alberta (Watson & Renney 1974); but these amounts are small compared to the estimated<br />

800000 ha in Montana, United States (Maddox 1979). The western Canadian distribution<br />

<strong>of</strong> the weed in 1977 is shown in Fig. 5.<br />

It is predicted that spotted knapweed will spread to dominate the herbaceous community<br />

in open uncultivated or lightly forested sites at the dry end <strong>of</strong> the Douglas fir zone<br />

in British Columbia and in the dark brown soils <strong>of</strong> Alberta and Saskatchewan. About 3.2<br />

million ha are vulnerable to invasion (Harris & Cranston 1979).<br />

Assuming that the two knapweeds spread to their predicted limits <strong>of</strong> approximately 10<br />

million ha. the loss in terms <strong>of</strong> dry native pasture species would be 2.5 million tonnes/year<br />

127


130 P.I-htrrisandJ. H. Myers<br />

Releases and Recoveries<br />

Urophora aflinis<br />

Frfld. (Diptera:<br />

TephrUldae)<br />

Urophora quadr1fBSCiata<br />

Mg.<br />

(Dlptera:<br />

Tephritidae)<br />

Cumm. The urediospores <strong>of</strong> P. jaeeae and P. eentaureae are similar in electron photomicrographs<br />

(Traquair el al. 1981), but are different from those <strong>of</strong> P. earlham; (Traquair<br />

1982 personal communication).<br />

It has recently been realized that because <strong>of</strong> different taxonomic treatments <strong>of</strong> spotted<br />

knapweed in Europe and North America, the problem species <strong>of</strong> British Columbia has<br />

been largely missed in the surveys done in Europe for possible biological control agents.<br />

Dostal (1976) recognizes 14 European species in the section Maculosae, whereas Moore<br />

& Frankton (1974) classified all the knapweeds in the section that have been introduced<br />

to North America as C. maeulosa. According to Dostal, C. maeulosa is diploid<br />

(2n= 18), whereas the problem knapweed <strong>of</strong> British Columbia is tetraploid (2n=36) and<br />

seems to be referable to C. biebersleinii D.C. C. biebersleinii is endemic to central and<br />

southeast Europe, but most <strong>of</strong> the survey was done on the more western species <strong>of</strong> C.<br />

valles;aea (D.C.) Jordan, C. maculosa Lam. s. SIr., and C. rhenana Bor.<br />

U. a/finis oviposits into immature flower heads <strong>of</strong> diffuse and spotted knapweed. The<br />

preferred head in diffuse knapweed is 5.5 to 7.5 mm long (Berube 1980). The larva<br />

stimulates gall formation from the receptacle tissues and feeds on enlarged and modified<br />

parenchyma cells within the gall. The number <strong>of</strong> galls that can develop in a<br />

head is proportional to the receptacle area, so the heads <strong>of</strong> spotted knapweed tend to<br />

have more galls than those <strong>of</strong> diffuse knapweed (Harris 1980a). The mature larva<br />

overwinters in the gall in the seed-head and its emergence in the spring is synchronized<br />

with the appearance <strong>of</strong> the first flower buds which are usually all attacked. In British<br />

Columbia, the partial second generation <strong>of</strong> U. affinis is too small to fully utilize the<br />

diffuse knapweed flower heads produced later in the summer (Roze 1981). However,<br />

spotted knapweed heads tend to be produced concurrently in early summer so that a<br />

smaller proportion escape attack.<br />

The releases <strong>of</strong> U. affinis are listed in Table 32. U. affinis established readily on both<br />

spotted and diffuse knapweed in British Columbia (Harris 1980a). It increased to a<br />

density <strong>of</strong> up to 3000 gallslml and spread widely throughout British Columbia to substantially<br />

reduce knapweed seed production. In Hastings County, Ontario, some galls<br />

were found in the year following release but heavy grazing destroyed most <strong>of</strong> the<br />

knapweed heads and the site has not been surveyed since. In Alberta, there was a<br />

change in policy which resulted in the treatment <strong>of</strong> all knapweed infestations with<br />

picloram. Flies became established on at least one site but we assume no colonies<br />

currently survive. The Quebec colonies released in 1980 were still present in 1981 and<br />

appear to be established (Watson 1981 personal communication).<br />

U. quadri/asciala is bivoltine and like U. affinis forms galls in the flower heads <strong>of</strong> a<br />

number <strong>of</strong> knapweed species. It emerges at the same time as U. affmis in the spring but<br />

oviposits into slightly more mature heads (7.5-9.5 mm in diffuse knapweed (Berube<br />

1980». The gall is fonned from the ovary wall tissues; several galls are nonnally found in<br />

a head and can develop in the same head as U. affinis. Both U. affmis and U.<br />

quadri/asciala increased rapidly after release, but at the original release sites U. quadrifasciala<br />

has since declined to low numbers, while U. affinis has stabilized at moderate<br />

levels <strong>of</strong> attack (Fig. 6). U. quadri/aseiata has spread more widely and rapidly than U.<br />

affinis and is now found on even small remote stands <strong>of</strong> knapweed in British Columbia.<br />

Berube (1980) suggested that U. quadrifaseiata has been less successful because heads


132 P. Harris and J. H. Myers<br />

Metzneria<br />

paucipuncteJJa<br />

Zell. (Lepidoptera:<br />

Gelechiidae)<br />

Table 29<br />

The life cycle <strong>of</strong> the moth on C. vaJlesiaca in the Swiss Rhone Valley (the source <strong>of</strong> the<br />

stock released in British Columbia) is reported by Englert (1971, 1972). The moth is<br />

univoltine and emerges at the end <strong>of</strong> May. It lays 60-100 eggs with a maximum <strong>of</strong>three eggs<br />

on an individual flower head base or adjacent stem. A single larva survives to develop in a<br />

head where it feeds principally on the achene while the seed coat is still s<strong>of</strong>t. The mature<br />

larva overwinters in the seed head which in spotted knapweed remains standing all winter<br />

having shed its seed in the fall. Up to nine seeds are eaten by a larva and in 83 attacked<br />

heads <strong>of</strong> C. vaJlesiaca, 95% <strong>of</strong> the viable seed was destroyed. Usually one-third to one-half<br />

<strong>of</strong> the heads were attacked even though the plants were widely scattered. Approximately<br />

20% <strong>of</strong> the eggs and 30-40% <strong>of</strong> the larvae were parasitized in Europe. Field records <strong>of</strong> the<br />

moth are restricted to knapweeds in the section MacuIosae, although in the laboratory it<br />

developed on C. diffusa.<br />

Adult M. paucipunctella were released (Table 32) in 1973 and 1974 at Castelgar airport<br />

and Westwold, British Columbia. M. paucipunctella has not been recovered from<br />

Castelgar; the number released was slightly less than at Westwold (73 in 1973 and 76 in<br />

1974) and in 1974, the flowering plants were mown shortly afterthe release was made. At<br />

Westwold, the initial establishment was tenuous with only four larvae found at the<br />

release point in August 1973. By the following year, larvae were found up to 20 m away<br />

(Table 29) and in 1976, they had spread a radius <strong>of</strong> at least 50 m. By 1979, the population<br />

was evenly distributed within a 50 m radius and by 1981, M. paucipunctella had spread at<br />

least 0.5 km.<br />

Establishment <strong>of</strong> M. paucipunctella Zell. on spotted knapweed, C. maculosa Lam., at<br />

Westwold, British Columbia.<br />

Year<br />

(fall)<br />

1973<br />

1974<br />

1976<br />

1977<br />

1978<br />

1979<br />

1980<br />

Radius Sampled<br />

(m)<br />

25<br />

20<br />

50<br />

30<br />

25<br />

50<br />

15<br />

No. Heads<br />

Examined<br />

4351<br />

2999<br />

485<br />

1443<br />

1719<br />

531<br />

443<br />

% heads with a M.<br />

paucipunctella larva<br />

0.1<br />

0.7<br />

2.0<br />

3.7<br />

27.5<br />

20.3<br />

37.7<br />

The population growth <strong>of</strong> M. paucipunctella has reflected the winter mortality <strong>of</strong><br />

larvae in the previous year. For example 68% <strong>of</strong> the larvae survived the winter <strong>of</strong><br />

1977-78 compared with only 28% in 1978-79. The reason for the mortality is not clear.<br />

The larvae can survive colder temperatures than they experience at Westwold; survival in<br />

the mild humid winter <strong>of</strong> Vancouver in 1977-78 was 82%, and it was 78% in the dry cold<br />

winter <strong>of</strong> Walachin, British Columbia. Possibly larvae are affected by a combination <strong>of</strong><br />

moisture and cold and the rate <strong>of</strong> onset <strong>of</strong> cold weather. Larvae absorb and lose moisture<br />

readily: 14 overwintering larvae dissected from knapweed heads and put on moist filter<br />

paper at 1000C in 3 days increased their weight by 50% (from 3.17±0.56 g to 4. 79± 1.13 g).<br />

After 16 days in a desiccator, they had returned to their original weight <strong>of</strong> 3.11±0.66 g.<br />

Less than 1% <strong>of</strong> the larvae are parasitized by an Elachertus sp. (Chalcidae) and some<br />

larvae are eaten by mice which feed on the seed heads particularly in the late summer.<br />

The percentage <strong>of</strong> seed destroyed by M. paucipunctella in <strong>Canada</strong> is much lower than<br />

the 95% reported by Englert (1971). However, C. biebersteinii heads at Westwold


Sphenoptera<br />

jugoslav/ca Obenb.<br />

(Coleoptera:<br />

Buprestidae)<br />

Table 30<br />

Evaluation <strong>of</strong> Control Attempts<br />

Centallrea diffllsa Lam. and C. maclllosa Lam. s. lat., 133<br />

contain an average 17 seeds compared with the 9-10 seeds for C. vallesiaca in the Swiss<br />

Rhone Valley. Only one M. paucipunctella larva develops in a head and destroys<br />

approximately four seeds or about one quarter <strong>of</strong> the production in a head at Westwold.<br />

However, at Westwold, M. paucipunctella does coexist in the same heads with U.<br />

quadrifasciata and there was no significant difference in the number <strong>of</strong> galls in heads<br />

with and without M. paucipunctella. Certainly U. quadrifasciata is adding to the seed<br />

destruction by M. paucipunctel/a but the moth larvae destroy U. affinis galls in the head.<br />

The biology and host specificity <strong>of</strong> S. jugoslavica were studied by Zw61fer (1976). The<br />

beetle is indigenous to the Balkans where it is found only on C. diffusa and C. jurineafolia.<br />

The adults feed on the knapweed foliage and do relatively minor damage. The female<br />

oviposits between the bases <strong>of</strong> tightly appressed rosette leaves, and if rosette growth<br />

occurs during the egg stage, the larva is unable to penetrate into the root crown. Thus the<br />

beetle is restricted to regions with a reliable summer drought during this stage. From the<br />

root crown a single larva bores into the root leaving a cylinder <strong>of</strong> cortex undamaged. The<br />

rosette continues to live, but in Europe <strong>of</strong>ten does not flower in the following year and so<br />

is subject to a second attack by S. jugoslavica. Up to 70% <strong>of</strong> the larvae in Europe are lost<br />

to parasitoids, predators, and competitors, but they were able to coexist with insects<br />

that feed on the outside <strong>of</strong> the knapweed roots. The beetle is univoltine and produces<br />

33-65 eggs/female.<br />

S. jugoslavica was released at Grandforks (51 adults) and White Lake, British<br />

Columbia (188 adults) in 1976 (Table 32). No survivors were found at Grandforks, but<br />

the colony at White Lake increased to infest 25-50"10 <strong>of</strong> the rosettes within a 250 m<br />

radius in 1981, and over half the flowering plants had been attacked (Table 30).<br />

Proportion <strong>of</strong> diffuse knapweed plants, C. diffusa L., attacked by Sphenoplera at the<br />

White Lake Observatory release site.<br />

Year Distance From Release Site Total·<br />

II 3m 6m 10 m 13 m 26 m 39 m 52 m 65 m<br />

Stem<br />

PlantslO.25 ml X ±S.D. n<br />

1977 .13 (40)··<br />

1978 .38 (52) .29 (17) .07 (41)<br />

1979 .39 (18) .SO (5) .29 (7) .05 (20)<br />

1980 .40 ( 5) .SO (34) .46 (13)<br />

1981 .59 (94)in 23 quadrates 0-250 m<br />

• Ungrazed area only<br />

•• Number <strong>of</strong> flowering plants<br />

10.11<br />

3.1<br />

.SO (IS) .52 (23) .57 (21) .29 (14) .33 (9) 6.5<br />

4.3<br />

Predictions <strong>of</strong> the eventual effects <strong>of</strong> the knapweed gall flies, U. affinis and U. quadri·<br />

fasciata, are conflicting. The immediate effects <strong>of</strong> the flies on the study sites at Kamloops<br />

have been a decline in seed production from around 25 000 to 1 SOO/m l and a reduction in<br />

the biomass <strong>of</strong> the weed, but the number <strong>of</strong> plants per unit area has remained the same.<br />

Obviously any effects <strong>of</strong> seed destruction would be masked for several years by the seed<br />

bank in the soil which, under dense stands <strong>of</strong> knapweed, is massive. Survivorship<br />

studies by Roze (1981) indicate that 1 500 seedslm l should be enough to maintain the<br />

7.6<br />

3.5<br />

4.2<br />

1.8<br />

11<br />

IS<br />

8<br />

23


134 P. Harris and J. H. Myers<br />

Table 31<br />

density <strong>of</strong> the weed, so no decline should be expected. On the other hand, Berube &<br />

Myers (1982) suggested that the knapweed would be controlled on sites where there<br />

was a high seedling loss either from drought or in moist sites from strong competition<br />

from other vegetation. In a later study, Myers & Berube (in press) found that the<br />

number <strong>of</strong> knapweed plants surviving to the flowering stage was directly proportional to<br />

the number <strong>of</strong> seedlings over a wide range <strong>of</strong> knapweed densities. If the results <strong>of</strong> this<br />

study are generally applicable, the flies should eventually achieve a large decline in<br />

knapweed density. The difference in the two predictions may be related to the study<br />

methods. ROle (1981) achieved a low knapweed density by removing seedlings from<br />

plots <strong>of</strong> almost pure knapweed. This had the effect <strong>of</strong> temporarily reducing the competition<br />

for the remaining knapweed. In the Myers & Berube (in press) study, the<br />

knapweed was always in competition: with itself at high densities and with grass at low<br />

densities.<br />

The effect <strong>of</strong> the beetle S. jugoslavica on diffuse knapweed was not studied in detail as<br />

until recently the colony was too small for destructive sampling. That the beetle tends to<br />

reduce seed production is evident from Table 31. Its effect may be much greater if it<br />

Effect <strong>of</strong> S. jugoslavica Obenb. on diffuse knapweed, C. diffusa L., seed production<br />

per flowering plant at White Lake, British Columbia<br />

Attacked Unattacked<br />

Year X ±SD n X ±SD n<br />

1977 18.6 13.4 5 20.4 14.8 28<br />

1978 69.1 45.9 26 73.5 55.7 82<br />

1979 38.9 33.9 10 66.5 80.4 16<br />

1980 31.6 24.4 25 54.3 52.5 27 P 0.05 U Test<br />

tends to prevent rosettes from bolting: approximately 20% <strong>of</strong> the rosettes in the fenced<br />

area had been attacked in the previous year, so they had passed at least two years<br />

without flowering. No previously attacked rosettes were found in the grazed area but the<br />

significance <strong>of</strong> this is not known, nor is the percentage <strong>of</strong> rosettes that normally take two<br />

years to flower. There was an indication that S. jugsolavica had a synergistic effect on<br />

seed reduction in plants stressed by drought or attacked by the seed-head gall flies. Both<br />

gall flies were present on the site, and in the autumn <strong>of</strong> 1981, there was an average <strong>of</strong><br />

1.63 ± 1.43 (S.D.) galls in the five distal heads on 23 plants; but whether a combination <strong>of</strong><br />

S. jugoslavica and the flies reduce seed production below the level needed for<br />

maintenance <strong>of</strong> the knapweed stand remains to be seen.<br />

The effect <strong>of</strong> M. paucipunctella on spotted knapweed was less than expected from<br />

European studies as a different species <strong>of</strong> knapweed appears to be involved in British<br />

Columbia. Also, the overwintering larvae at Westwold, British Columbia, have suffered<br />

a periodic large and unexplained winter mortality that has prevented the rapid increase<br />

and spread <strong>of</strong> the species. However, without releases in other knapweed stands, it<br />

should not be assumed that this mortality occurs throughout the spotted knapweed<br />

region <strong>of</strong> North America. The ability <strong>of</strong> M. paucipunctel/a or the seed fly U.<br />

quadrifasciata to survive in the colder parts <strong>of</strong> the Canadian prairies is doubtful. In<br />

the winter <strong>of</strong> 1981, spotted knapweed heads containing overwintering larvae <strong>of</strong> the three<br />

seed-head insects were tied to a stake at normal plant height at Regina. Only U. affinis<br />

larvae were alive in the spring. If large stands <strong>of</strong> the weed became established on the<br />

prairies, it might be possible to select cold hardy strains <strong>of</strong> the insects but in the meantime


Table 32<br />

Recommendations<br />

Centaurea diffusa Lam. and C. maculosa Lam. s. lat., 135<br />

the safest course <strong>of</strong> action is to chemically treat stands <strong>of</strong> the weed as they appear and are<br />

still small.<br />

Auld et al. (1979) pointed out that the best control strategy against a weed on either a<br />

farm or regional basis depends on the rate at which it spreads. The faster the rate, the<br />

greater the necessity to treat the source area while for slowly spreading weeds, containment<br />

is the most economic course <strong>of</strong> action. The seed reduction achieved by the<br />

biological control agents already established in British Columbia should have decreased<br />

the rate <strong>of</strong> knapweed invasion from headlands onto tame pasture as well as the spread to<br />

new regions within western <strong>Canada</strong>. Unfortunately, the benefits from this reduced rate<br />

<strong>of</strong> spread will not be reflected in the chemical control programme until this is done on a<br />

strictly costlbenefit basis. To reduce knapweed with biological control agents to the<br />

point where the publicly funded chemical control programme becomes clearly unnecessary<br />

will require the establishment <strong>of</strong> more biological control agents on knapweed in<br />

<strong>Canada</strong>.<br />

Open releases and recoveries <strong>of</strong> insects against Centaurea diffusa L. and C. maculosa<br />

Lam. s. lat.<br />

Control Agent Province Released Year Origin Number Year <strong>of</strong> Recovery<br />

C. I1Ulcu/o5a<br />

Urophora affinis Ontario 1970 France 694 1971<br />

British Columbia 1970 France 297 1971<br />

Ontario 1971 France 97<br />

British Columbia 1971 France 377 1972<br />

British Columbia 1972 British Columbia 1472<br />

British Columbia 1974 British Columbia 550<br />

Alberta 1976 British Columbia 200 Site treated with herbicide<br />

Quebec 1979 British Columbia 950<br />

Quebec 1980 British Columbia 345<br />

Urophora quadrifa5c",ra Quebec 1979 British Columbia 950<br />

Merzneria paucipuncrel/a British Columbia 1973 Switzerland 197 1974<br />

British Columbia 1974 Switzerland 177 1975<br />

British Columbia 1979 British Columbia 180 Overwintering test<br />

C. diffusa<br />

Urophora affinis British Columbia 1970 France 284 1971<br />

British Columbia 1971 France 209 1972<br />

British Columbia 1972 USSR 797 1973<br />

Alberta 1977 British Columbia 681 Site treated with herbicide<br />

Quebec 1980 British Columbia<br />

Urophora quadrifa5ciara British Columbia 1972 USSR 50 1973<br />

Sphenoprera jugol/avica British Columbia 1976 Greece 239 1977<br />

(1) The effects <strong>of</strong> S. jugoslavica on diffuse knapweed in British Columbia should be<br />

determined. If the beetle adds to the seed reduction that is achieved by the seed-head<br />

gall flies. it should be established throughout the diffuse knapweed region.<br />

(2) The susceptibility <strong>of</strong> M. paucipunctella to winter mortality should be detennined for<br />

a number <strong>of</strong> places within the spotted knapweed region and the moth established in<br />

favorable sites.<br />

(3) Additional species <strong>of</strong> biological control agents should be established on both<br />

spotted and diffuse knapweed in <strong>Canada</strong> as soon as they can be screened and approved


136 P. Harris and J. H. Myers<br />

Acknowledgements<br />

Literature Cited<br />

for release. In particular. studies should be completed on the root-feeding moth Agapeta<br />

zoegana for spotted knapweed. and Pelochrisla medullana for diffuse and spotted<br />

knapweed.<br />

(4) The possibility <strong>of</strong> using or enhancing pathogens already on knapweed in North<br />

America. particularly SclerOlina sclerotiorum and Puccinia jaceae should be explored.<br />

Although S. scleroliorum is a major pest <strong>of</strong> vegetable crops. its presence on knapweed<br />

infested rangeland should not aggravate the problem on cultivated land.<br />

Aspects <strong>of</strong> the biological control <strong>of</strong> knapweed have been the subject <strong>of</strong> Ph.D. studies at<br />

the University <strong>of</strong> British Columbia (U .B.C.) by Liga Roze and Peter Morrison. <strong>of</strong> undergraduate<br />

or technical studies. from U.B.C. by Barbara Rawlek. Woody Bennet. Sandy<br />

Ockenden and from Regina by Julie Soroka. Their work and ideas have been used extensively<br />

in preparing this review. Thanks are also due to Peter Morrison and Julie Soroka<br />

for reviewing and suggesting changes in the manuscript.<br />

Auld. B.A.; Coote. B.G.; Menz. K.M. (1979) Dynamics <strong>of</strong> plant spread in relation to weed control. ProCt'edings <strong>of</strong> 'he 7,h Asian Pacific<br />

Weed Science Conference. 399-402.<br />

Berube. D. E. (1980) Interspecific competition between Urophora affinis and U. quadrifascia,a (Diptera: Tephritidae) for ovipositional sites<br />

on diffuse knapweed (Centaurfa difflLta Compositae). ZeiLtclirif' far angewand'e Entonwlogie 90.<br />

299-306.<br />

Berube. D.E.; Myers. J.H. (1982) Suppression <strong>of</strong> knapweed invasion by crested wheatgrass in the dry interior <strong>of</strong> British Columbia. Journal <strong>of</strong><br />

Range Matlagemen' 35. 459-461.<br />

Cummins. G.B. (1977) Nomenclature changes and new species in the Uredinales. Myco,axotl 5. 398-408.<br />

Dostal. J. (1976) Cen'aurea L. p. 254-301. In: Flora europaea. vol 4. Cambridge University Press. 505 pp.<br />

Englen. W. (1971) Me'Ztleria paucipunc,ella Zell. (Ge1echiidae. Lepidoptera): a potential insect for the biological control <strong>of</strong> Cenlaurea<br />

s,oebe L. in <strong>Canada</strong>. Commotlweal,h Ins'i'u'e <strong>of</strong> Biological Cotllrol Progress Repor' 28. 12 pp.<br />

Englen. W.O. (1972) Revision der Gallung Me,zneria Zeller (Lepidoptera. Gc1echiidae) mit Beitriigen zur Biologic der Anen. Zeitscllrif' far<br />

angewand'e Entomologie 75.381-421.<br />

Fletcher. R.A.; Renne)" A.J. (1963) A growth inhibitor found in Centaurea spp. Canadian I(}urnal (}f Plant Science 43.475-481.<br />

Groh. H. (1940) Turkestan alfalfa as a medium <strong>of</strong> weed introduction. Science in Agricul,ure 21.36-43.<br />

Groh. H. (1943) Canadian weed survey. 2nd Annual Repon <strong>of</strong> the Canadian Department <strong>of</strong> AgriCUlture. 74 pp.<br />

Harris. P. (1979) Cost <strong>of</strong> biological control <strong>of</strong> weeds by insects in <strong>Canada</strong>. Weed ScienCt' 27. 242-250.<br />

Harris. P. (19800) Establishment <strong>of</strong> UropllOra affitlis Frnd. and U. quadrifascia,a (Meig.) in <strong>Canada</strong> for the biological control <strong>of</strong> diffuse and<br />

spOiled knapweed. Zeitschrif' fiir atlgewand'e En'(}m(}logie 89. 504-514.<br />

Harris. P. (1980b) Effects <strong>of</strong> Urophora affinis Frnd. and U. quadrifascia,a (Meig.) (Diptera: Tephritidae) on Om'al/rea diffusa Lam. and C.<br />

maclliosa Lam. (Compositae). Zei'schrif' far angewatld'e En'omologie 90. 190-210.<br />

Harris. P.; Cranston. R. (1979) An economic evaluation <strong>of</strong> control meusures for diffuse and spOiled knapweed in western <strong>Canada</strong>. Canadian<br />

10llmal <strong>of</strong> I'latl' Scitnce 59. 375-382.<br />

Howell, J.T. (1959) Distributional data on weedy thistles in western North America. Leafle, <strong>of</strong> Wes'ern Bo'any 9.17-29.<br />

Jeffrey. C. (1967) Notes on Compositae: 111. The Cynareae in cast tropical Africa. Kew Bulle,in 22, 107-140.<br />

Maddox, D.M. (1979) The knapweeds: their economics and biological control in the western States. U.S.A. Rangelatlds I. 139-143.<br />

Moore. R.J.; Frankton. C. (1974) The thistles <strong>of</strong> <strong>Canada</strong>. Calladiotl Drpar'menl <strong>of</strong> Agricul,ure Motlonograph 10. Ottawa. 111 pp.<br />

Myers. J. H.; Berube. D. E. (in press) Diffuse knapweed invasion into rangeland in the dry interior <strong>of</strong> British Columbia. Canadian Journal <strong>of</strong><br />

Plan' Science.<br />

Myers. J.H.; Harris, P. (1980) Distribution <strong>of</strong> Uroplwra gulls in nower heads <strong>of</strong> liffuse and spotted knapweed in British Columbia. JOl/rnal <strong>of</strong><br />

Applied Ecol(}gy 17.359-367.<br />

Renney. A.J. (1959) Centaurea spp. infestation in British Columbia. Proceedinll$ <strong>of</strong> a Joint Meeting <strong>of</strong> the North Central Weed Control<br />

Conference 16 and the Western <strong>Canada</strong> Weed Control Conference 10. 18-19.<br />

Roze. L. (1981) The biological control <strong>of</strong> Centaurea diffll.ta Lam. and C. maclIl(}.ra Lam. by Uropllora affitlis Frauenfeld and U.<br />

quadriflLrcia'a Meigel! (Diptera: Tephritidae). Ph.D. thesis. University <strong>of</strong> British Columbia.<br />

Savile. D.B.O. (1970a) Some Eurasian Puccinia species allacking Cardueae. Canadian 10llmal <strong>of</strong> BOIony 48. 1553-1566.<br />

Savile. D.B.O. (1970b) Autoecious Pllccinia species al1acking Cardueae in North America. Canadian loumal <strong>of</strong> Bo'any 48. 1567-1584.


Celltaltrea diffusa Lam. and C. macltlosa Lam. s. lat., 137<br />

Traquair, J.A.; Kokko, E.G.; Harris, P. (1981) Urcdiniosporc morphology <strong>of</strong> two rust fungi on Cellllllm:ll (knapweed). Almrllcls ollile<br />

Cll/ladiall 80latlical Assoeiatioll Amlllaf Meelitlg 1981.<br />

Watson, A.K.; Alkoury, I. (1981) Response <strong>of</strong> safflower cultivars to Pllceillia jaceae collected from diffuse knapweed in eastern Europe.<br />

Proceeditlgs olllle 5111 Itilemuliolllif SymptJSillt1l 1m Ille Biological Cotllml <strong>of</strong> Weeds, 301-305.<br />

Watson, A.K.; Copeman, R.J.; Relllley A.J. (1974) A first record <strong>of</strong> Scleroliml sclerotio",,,, and Microsplllleropsis cetllllllreac on CCtllllllrell<br />

diffllsa. Call1ldiatl Jmmlllfl11 Botatly 52. 2639-2640.<br />

Watson, A.K.; Renney, AJ. (1974) The biology <strong>of</strong> Canadian weeds, 6. Celllmlreu diffiLm ,lIId C. Itrllcillosa. Camldiutl JOllmal <strong>of</strong> PfulI/<br />

SciC/lce 54,687-701.<br />

Watson, A. K.; Schroeder, D.; Alkoury, I. (1981) Collection (If Pllccitlill specics [rom diffusc knapweed in eastern Europe. Call1uli,m JOllmaf<br />

01 Plam Palhology 3.6-8.<br />

Zwolfer H. (1976) Investigations on SplJetloplera (Cllilo.rletlla) jllgosfal·ica Obenb. (Col. Bupreslidac), a possiblc bioconlrol agent <strong>of</strong> the<br />

weed CetililllTeU diffllsil Lalli. (Colllpositac) in <strong>Canada</strong>. ZcilSc/rrifr liir Iltlgewalldre EII/oltralogie SO,<br />

170-190.


Blank Page<br />

138


Pest Status<br />

Background<br />

Releases and Recoveries<br />

Altics csrduorum<br />

Guer. (Coleoptera:<br />

Chrysomelidae)<br />

Chapter 32<br />

Cirsium arvense (L.) Scop., <strong>Canada</strong><br />

Thistle (Compositae)<br />

D.P. PESCHKEN<br />

New data have been published on the pest status <strong>of</strong> <strong>Canada</strong> thistle, Cirsium arvense (L.)<br />

Scop., since the previous ten year review (Peschken 1971). Thomas (1980) summarized<br />

the abundance <strong>of</strong> <strong>Canada</strong> thistle on cultivated land in <strong>Canada</strong>. This is a particularly<br />

serious weed in the three Prairie Provinces, especially in the black soil zone and in<br />

Manitoba. In this province it ranked fifth and sixth in abundance in 1978 and 1979<br />

respectively; ninth and twelfth in Saskatchewan from 1976 to 1979; and on the average<br />

thirteenth in Alberta from 1973 to 1979. Of their most troublesome weeds, farmers in<br />

Alberta, Saskatchewan, and Manitoba ranked <strong>Canada</strong> thistle second, fifth, and third<br />

respectively, i.e. considerably higher than the rank based on relative abundance (Thomas<br />

1980). Peschken et al. (1980) estimated the loss in yield <strong>of</strong> wheat in Saskatchewan based<br />

on the sampling estimate <strong>of</strong> the density <strong>of</strong> <strong>Canada</strong> thistle in wheat fields (Thomas 1976)<br />

and estimates <strong>of</strong> the reduction in yield <strong>of</strong> wheat at various density levels <strong>of</strong> <strong>Canada</strong><br />

thistle. Assuming the 10 year average yield <strong>of</strong> 1607 kg per ha and a price <strong>of</strong>$14.7 per 100<br />

kg, Saskatchewan farmers lost 24 300 t <strong>of</strong> wheat or $3 600 000 in the reduction <strong>of</strong> yield <strong>of</strong><br />

wheat alone due to <strong>Canada</strong> thistle in 1976. The costs <strong>of</strong> chemical and cultural control <strong>of</strong><br />

<strong>Canada</strong> thistle are high. Thus the total cost <strong>of</strong> <strong>Canada</strong> thistle in terms <strong>of</strong> cost <strong>of</strong> control<br />

and losses to all crops is a multiple <strong>of</strong> the $3 600 000 estimated for wheat.<br />

Natural enemies <strong>of</strong> <strong>Canada</strong> thistle and the initial stages <strong>of</strong> the biological control work<br />

were discussed in the previous review (Peschken 1971). Work on the biological control<br />

<strong>of</strong> <strong>Canada</strong> thistle has also been done in the United States with Altica carduorum Ouer.<br />

(Coleoptera: Chrysomelidae), Urophora cardui (Diptera: Tephritidae), and Ceutorhynchus<br />

Iitura (Coleoptera: Curculionidae) (Andres 1980, Schaber et al. 1975, Story<br />

1980). The latter species is established in Idaho, South Dakota, and Montana (Julien,<br />

1982). Baker et al. (1972) reported on releases <strong>of</strong> A. carduorum against <strong>Canada</strong> thistle in<br />

England. Ward & Pienkowski (1978a, 1978b), working in the state <strong>of</strong> Virginia, United<br />

States, investigated the biology and mortality <strong>of</strong> Cassida rubiginosa Muell. (Coleoptera:<br />

Chrysomelidae) which has been accidentally introduced into eastern North America.<br />

One insect was screened but not released. The only confirmed host plant <strong>of</strong> Tingis<br />

ampliata H.-S., (Heteroptera: Tingidae) in its native habitat in Europe is C. arvense.<br />

However, in the laboratory it developed fertile eggs on several other plants. including<br />

the cultivated globe artichoke (Cynara scolymus L.) and safflower (Carthamus tinctorius<br />

L.). Therefore, T. ampliata was not recommended for release (Peschken 1977).<br />

(a) Ecology<br />

Baker et al. (1972) showed that the development <strong>of</strong>the immature stages <strong>of</strong> A. carduorultl<br />

is slightly faster at 100% relative humidity (r.h.) than at 97%. At an average temperature<br />

139


140 D. P. Peschken<br />

Ceutorhynchus<br />

lItura (F.)<br />

(Coleoptera:<br />

Curcullonidae)<br />

<strong>of</strong> 15°C, development was not completed. while at 22.5°C to 25°C. it was completed in 33-<br />

38 days. Thus high temperatures and humidities are needed for an optimal rate <strong>of</strong><br />

development. This agrees with the coastal Mediterranean and Atlantic native distribution<br />

<strong>of</strong> the beetles (Peschken 1971).<br />

(b) Releases<br />

Releases up to and including 1968 were reported by Peschken (1971). The colony at<br />

Lacombe, Alberta survived until 1971. In addition, 170 were released in 1969 at Fon<br />

Vermilion, Albena (58°28' N), which is the most northern and coolest <strong>of</strong> all release sites<br />

<strong>of</strong> A. carduorum in <strong>Canada</strong> (Table 33). In 1970, 1018 beetles were released at Essex,<br />

Ontario (42° 22' N), which is the most southern and warmest <strong>of</strong> all release sites. Both<br />

releases were made early in the spring which prevented overdispersal in high<br />

temperatures (Peschken 1977). Predation <strong>of</strong> eggs and larvae was severe at Essex and less<br />

severe at Fort Vermilion. In the fall <strong>of</strong> the respective release years, 18 adults were found<br />

at Fon Vermilion and only one at Essex, although six times as many beetles had been<br />

released at the latter site. At Essex, the carabid beetle Lebia viridis Say, was implicated as<br />

a main predator <strong>of</strong> eggs and larvae. In feeding tests it consumed 0.7 eggs and larvae per<br />

hour (Peschken 1971). No A. carduorum were found in subsequent years.<br />

(a) Ecology<br />

The ecology <strong>of</strong> this stem-mining weevil was summarized by Peschken (1971). Since then<br />

Peschken & Beecher (1973) and Peschken & Wilkinson (1981) have published additional<br />

information. The weevil larvae feed on the parenchyma tissue <strong>of</strong> the stem and avoid the<br />

vascular bundles. In <strong>Canada</strong>, the females oviposit from early May until the beginning <strong>of</strong><br />

June. The new adults copulate and feed heavily during sunny, late summer and early fall<br />

days ",ith temperatures over 18°C, but do not oviposit. At this time, males outnumber<br />

females 2: 1.<br />

In spring, females outnumber the males 1.6: 1. In the laboratory females laid an average<br />

<strong>of</strong> 122 eggs in temperatures varying from 17°C to 27°C during a 16.5 hour day and 7°C to<br />

17°C during a 7.5 hour night. An average <strong>of</strong> 80% <strong>of</strong> the eggs hatched.<br />

(b) Releases<br />

A total <strong>of</strong> 1188 weevils was released in five provinces at 11 release sites (Table 33).<br />

The insect became established at nine sites in varying climates in four provinces <strong>of</strong><br />

<strong>Canada</strong>. It spread slowly, 2.9 km in 12 years in Ontario, 90 m and 75 m in six and four<br />

years respectively at two release sites in Saskatchewan. The length <strong>of</strong> mines varied on<br />

the different release sites from 5 to 23 cm. The short, 5 cm, mines were found at<br />

Lacombe, Albena where almost all <strong>of</strong> the thistles were smooth-leaved, resembling<br />

those <strong>of</strong> C. arvense var. integrifolium Wimm. & Grab. The weevil preferred dense<br />

patches <strong>of</strong> thistle.<br />

Mined shoots were from 1.09 to 1.96 times taller than unmined ones in the fall<br />

(Peschken & Wilkinson 1981). This is explained as follows: early emerged, unattacked<br />

thistle rosettes grow into taller shoots by fall than later emerged ones (Peschken &<br />

Wilkinson 1981). The weevil oviposits into the early emerged rosettes and it is


Urophora cardui<br />

(L.) (Diptera:<br />

Tephritidae)<br />

Cirsium arvense (L.) Scop.. 141<br />

assumed that larval mining does not prevent the early emerged rosettes from growing<br />

into taller shoots. This also means that damage to the thistles is light.<br />

The development <strong>of</strong> the weevils and thistles was monitored in detail at one release<br />

site in Ontario. The weevil was released on a dense patch <strong>of</strong> <strong>Canada</strong> thistle, and<br />

eventually it spread to other more or less contiguous thistle patches on the same<br />

permanent pasture covering about 1 ha. On the original release patch, the level <strong>of</strong><br />

mined stems peaked at 72% in the fifth growing season with an average <strong>of</strong> 2.75 larvae<br />

per stem, and thistle density subsequently diminished to zero (Peschken & Beecher<br />

1973, Peschken & Wilkinson 1981). However, on another patch in the same pasture,<br />

thistle density did not diminish although 77-91 % <strong>of</strong> the stems were mined for four<br />

years, while on a third thistle density declined to zero per m 1 with only up to 2% <strong>of</strong><br />

the stems being mined. Thus no consistent correlation between an increase in the<br />

weevil population and a decline <strong>of</strong> the thistles was observed. The rust Puccinia<br />

punctiformis (Str.) Roh!. and the thistle feeding beetles Cleonus piger Scop. (Coleoptera:<br />

Curculionidae) and Cassida rubiginosa Muell. (Coleoptera: Chrysomelidae) were also<br />

present on this release pasture, and dense thistles tend to decrease to scattered shoots<br />

behind an advancing front (Amor & Harris 1975). It is assumed that the decline <strong>of</strong><br />

thistles on some <strong>of</strong> the patchcs was caused by one or a combination <strong>of</strong> these factors.<br />

Similarly, the fluctuating thistle densities at release sites in western <strong>Canada</strong> could not<br />

be associated with levels <strong>of</strong> weevil infestation.<br />

(a) Ecology<br />

This gall fly occurs from France in the west (ZwOlfer 1967) to near the Crimea<br />

(Oirlbeck & Oirlbeck 1964) and Siberia in (he east (Leclercq 1967), and from Sweden<br />

in the north (Zetterstedt 1847) to the Mediterranean in the south (Zwolfer 1967).<br />

Areas in half shade seem to be preferred over those in full sun (ZwOlfer 1967).<br />

The fly is monophagous on <strong>Canada</strong> thistle and oviposits into the vegetative buds <strong>of</strong><br />

the main or side shoots. In Saskatchewan, the oviposition period extends from the end<br />

<strong>of</strong> May to the third week in July. The larvae overwinter and pupate in spring inside a<br />

multi-locular lignified gall. There is a strong correlation in the number <strong>of</strong> larvae and the<br />

size <strong>of</strong> field collected or laboratory produced galls (Pesch ken et al. 1982).<br />

(b) Releases<br />

A total <strong>of</strong> 4932 flies and 290 galls was released in 14 locations in the four western<br />

provinces <strong>of</strong> <strong>Canada</strong> and in 10 locations in Ontario, Quebec, and New Brunswick (Table<br />

33). In western <strong>Canada</strong>, galls were usually produced in the year <strong>of</strong> the release, and on<br />

two sites it survived for two and three years respectively. However, the fly died out in<br />

western <strong>Canada</strong> except for a very small colony at Camrose, Alberta.<br />

In contrast, U. cardu; is established at five locations in Ontario, Quebec, and New<br />

Brunswick. Near Sussex, New Brunswick, the fly had been released at two locations,<br />

4.5 km apart. The population <strong>of</strong> one release had spread over 1000 ha by 1980, but only<br />

6% <strong>of</strong> the shoots were galled in anyone thistle patch. By 1981, populations <strong>of</strong> the two<br />

release sites had merged and spread over 3000 ha. On some thistle patches up to seven<br />

galls were found on one thistle shoot and there was an average <strong>of</strong> three on the attacked<br />

shoots (O.B. Finnamore. 1981. personal communication). In 1979. 13% <strong>of</strong> a sample <strong>of</strong><br />

larvae was diagnosed to contain Nosema sp. (Microsporidia) spores and the percentage<br />

was higher in dead than in live larvae. Three percent <strong>of</strong> the larvae were parasitized by<br />

Habrocyrus elevatus Walker (Hymenoptera: Pteromalidae). In 1980, only 3% <strong>of</strong> the


142 D. P. Pcschken<br />

Lema cyanella<br />

(L.) (Coleoptera:<br />

Chrysomelidae)<br />

larvae contained Nosema spores and 1.3% were parasitized by an as yet unidentified<br />

hymenopterous parasite. In Saskatchewan and Alberta, the mortality was higher: 35%<br />

<strong>of</strong> a total <strong>of</strong> 402 larvae collected in the spring <strong>of</strong> 1975, 1976, and 1977 was dead.<br />

Comparing only those thistles that emerged during the oviposition period <strong>of</strong> the fly,<br />

shoots with galls on the main shoot were 57% shorter, and those with sideshoot galls<br />

were 10% shorter than ungalled shoots in Quebec when growing on an uncultivated site<br />

with competing vegetation. However, in Saskatchewan even an average <strong>of</strong> 13 galls per<br />

shoot did not reduce the height, dry weight, number <strong>of</strong> seed heads or the dry weight <strong>of</strong><br />

new shoots produced by thistles growing in the absence <strong>of</strong> competition on heavy fertile<br />

soil with sufficient moisture (Peschken et al. 1982).<br />

(a) Ecology<br />

This leaf-feeder is widely distributed in Europe and A!>ia between latitudes 35°N and<br />

65°N and occurs about three-quarters <strong>of</strong> the range <strong>of</strong> C. arvense (Peschken & Johnson<br />

1979). It is rare in Germany, Austria, Czechoslovakia, and Poland and is more frequent<br />

in western Europe (Zwolfer & Pattullo 1970, Winiarska 1973). Collection records<br />

indicate that moist habitats are preferred (Winiarska 1973, ZwOlfer & Pattullo 1970,<br />

Lopatin 1960, Peschken unpublished).<br />

The adult beetles hibernate and in Europe they appear on the thistle rosettes at about<br />

the end <strong>of</strong> April. They oviposit in May and the new generation <strong>of</strong> adults can be observed<br />

from the end <strong>of</strong> June until August. There is one generation per year.<br />

(b) Host specificity<br />

The only confirmed breeding host in Europe is C. arvense (Zwolfer & Pattulo 1970,<br />

Peschken & Johnson 1979). In the screening tests for host specificity, it did not feed on<br />

any cultivated plants but on several Cirsium spp. native to North America (Peschken &<br />

Johnson 1979). One <strong>of</strong> these, Cirsium drummondiiT. & G., was even preferred in favour<br />

<strong>of</strong> <strong>Canada</strong> thistle in choice tests in the laboratory. In a field cage, however, the females<br />

laid most eggs on those thistles which were most prevalent in terms <strong>of</strong> biomass <strong>of</strong> leaves<br />

and stems (Peschken unpublished).<br />

(c) Releases<br />

The releases listed on Table 33 were made for the purpose <strong>of</strong> testing the host preference<br />

<strong>of</strong> L. cyanella in the more natural environment <strong>of</strong> a field cage, not for establishment in<br />

the field. In the course <strong>of</strong> these field cage studies, 4 out <strong>of</strong> 28 beetles survived the 1978n9<br />

winter. In subsequent years the beetles were collected in the fall and overwintered in the<br />

laboratory. However, another adult was collected in the field in the spring <strong>of</strong> 1981. Thus<br />

L. cyan ella can survive the winter in Saskatchewan. No larvae were found in 1981. The<br />

release <strong>of</strong> L. cyanella with the intent <strong>of</strong> establishment has not yet been approved out <strong>of</strong><br />

concern for the Cirsium species <strong>of</strong> our native North American flora.


Table 33 Open releases and recoveries <strong>of</strong> insects against Cirsium arvense (L.) Scop.<br />

Cirsium arvense (L.) Scop.. 143<br />

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

Species and Province Origin Year Number Recovery<br />

Altica carduorum (Guerin-Meneville)<br />

British Columbia Switzerland· 1969 5534<br />

Alberta Switzerland· 1969 170 1971<br />

Ontario Switzerland • 1970 1018<br />

France 1970 345<br />

TOTAL 7067<br />

Ceutorhynchus litura (F.)<br />

British Columbia Germany, Switzerland 1975 69 1976-80<br />

France, Italy<br />

Alberta Germany, Switzerland 1975 57 1976-80<br />

France, Italy<br />

Indian Head, Saskatchewan 1978 280 1979-80<br />

Saskatchewan Switzerland· 1973 70 1974-80<br />

Switzerland· 1974 30 1975-80<br />

Germany, Switzerland 1975 41 1976-80<br />

France, Italy<br />

Switzerland· 1976 56<br />

Switzerland • 1978 25 1979<br />

Indian Head, Saskatchewan 1979 228 1980<br />

Ontario Switzerland·· 1966 56<br />

New Brunswick Indian Head, Saskatchewan 1978 276 1981<br />

TOTAL 1188<br />

Urophora cardui (L.)<br />

British Columbia Germany 1974 274<br />

Germany 1975 98<br />

France. Austria 1976 199<br />

plus lab. reared<br />

from French stock<br />

Alberta Germany 1975 100 1976<br />

Austria, France 1976 472<br />

plus lab. reared<br />

from French stock<br />

Lab. reared from stock 1977 406 1978-81<br />

collected in <strong>Canada</strong><br />

Saskatchewan Germany 1974 625<br />

Germany, plus lab. 1975 274 1976-80<br />

reared from German stock<br />

France, Austria 1976 915 1977-78<br />

plus lab. reared<br />

from French stock<br />

Lab. reared from stock 1977 392<br />

collected in <strong>Canada</strong><br />

Ontario Germany 1975 81 1976-80<br />

France, Austria 1976 52 1977-80<br />

plus lab. reared<br />

from French stock


144 D. P. Peschken<br />

Table 33 continued<br />

Evaluation <strong>of</strong> Control Attempts<br />

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

Species and Province Origin Year Number Recovery<br />

Quebec Germany 1975 96<br />

France, Austria 1976 451 1977-79<br />

plus lab. reared<br />

from French stock<br />

Lab. reared from stock 1977 99<br />

collected in <strong>Canada</strong><br />

Field collected 1979 290 1981<br />

at Compton, Quebec galls<br />

New Brunswick France, Austria 1976 191 1977-82<br />

plus lab. reared<br />

from French stock<br />

Lab. reared from stock 1977 207 1978-82<br />

collected in <strong>Canada</strong><br />

TOTAL 4932 flies<br />

plus 290<br />

galls<br />

Lema cyanella (L.)<br />

Switzerland· 1978 28 1979<br />

1979 68<br />

1980 35 1981<br />

1982 31 1982<br />

TOTAL 162<br />

• These were laboratory reared beetles from stock collected in Switzerland .<br />

.. Not reported in Peschken 1971.<br />

The cool climate combined with predation prevented establishment <strong>of</strong> A. carduorum.<br />

The weevil, C. litura, is established in four provinces and probably can be colonized in<br />

all agricultural areas <strong>of</strong> <strong>Canada</strong>. However, damage to <strong>Canada</strong> thistle by larval mining is<br />

small. Furthermore, <strong>Canada</strong> thistle is a major weed in the grain belt <strong>of</strong> the three Prairie<br />

Provinces, where destruction <strong>of</strong> eggs and larvae by cultivation <strong>of</strong> thistles along field<br />

margins prevents the buildup <strong>of</strong> dense populations. U. cardu; is not established in<br />

western <strong>Canada</strong> (except for a small colony at Cam rose , Alberta), but it can probably be<br />

established anywhere in the agricultural areas <strong>of</strong> eastern <strong>Canada</strong>. The galls do cause<br />

noticeable stress to thistles growing in competition with other vegetation (Pesehken el<br />

al. 1982). Overdispersal prevented the buildup <strong>of</strong> dense populations and reduced<br />

effectiveness so far. It is not known why U. cardu; did not become established in western<br />

<strong>Canada</strong>. The 35% rate <strong>of</strong> mor.tality alone does not explain the failure <strong>of</strong> the release<br />

colonies, many <strong>of</strong> which bred well in the year <strong>of</strong> their release. Possibly severe spring<br />

frosts during the pupation period were a problem. Most release stock originated from the<br />

upper Rhine Valley which is considerably warmer and moister than the Prairie Provinces.<br />

Lalonde & Shortbouse (1982) suggest that inadequate moisture on the release sites in<br />

western <strong>Canada</strong> prevented the breakdown <strong>of</strong> the callus plug in the exit channel out <strong>of</strong> the


146 D. P. Peschken<br />

Ferdinandsen. C. (1923) Biologiske Undersogelser over Tidselrust (puccinia sual'eolens (Pers. Rostr.». Beretning om Nordi.rke jordbrllgsforskeres<br />

forenings Kongres 5-8, 475-487.<br />

Julien, M.H. (Ed.) (1982) Biological control <strong>of</strong> wceds: a world cataloguc <strong>of</strong> agents and their target weeds. Commonwealth Agricultural<br />

Bureaux. 108 pp.<br />

Lalonde. R.O.; Shorthousc, J.D. (1982) Exit strategy <strong>of</strong> Urophora cardui (Diptera: Tephritidae) from its gall on <strong>Canada</strong> thistle. Canadian<br />

Entomologist 114. 873-878.<br />

Leclercq. M. (1967) Contribution it I'etude des Trypetidae (Diptera) palearctiques et de leurs relations avec les vegetaux. Bulletin des<br />

Recherches Agronomiques de Gembloux (N.S.) 2, 64-105.<br />

Lopatin, I.K. (1960) Data on the fauna and ecology <strong>of</strong> leaf-beetles (Coleoptera: Chrysomelidae) on the southern rear bank <strong>of</strong> the Dnieper<br />

River. Entomologicheskoe Obozrenie 39(3).447-457.<br />

Peschken. D.P. (1971) Cinium arvense (L.) Scop .• <strong>Canada</strong> Thistle. In: Biological control programmes against insccts and weeds in <strong>Canada</strong><br />

1959-1968. Commonwealth InstitUle <strong>of</strong> Biological Control Technical Communication 4. 79-83.<br />

Peschken, D.P. (1977) Host specificity <strong>of</strong> Tingis ampliata (Tingidae: Heteroptera): a candidate for the biological control <strong>of</strong> <strong>Canada</strong> thistle<br />

(Cirsium arvense). Canadian Entomologist 109.669-674.<br />

Peschken. D.P.; Beecher. R.W. (1973) Ceutorhynchus lilllra (Coleoptera: Curculionidae): biology and first releases for biological control<br />

<strong>of</strong> the weed <strong>Canada</strong> thistle (Cirsium arvense) in Ontario. <strong>Canada</strong>. Canadian Entomologist 105.<br />

1489-1494.<br />

Peschken, D.P.; Finnamore D.B.; Watson, A.K. (1982) Biocontrol <strong>of</strong> the weed <strong>Canada</strong> thistle (Cirsium arvense): releases and<br />

development <strong>of</strong> the gall fly Urophora cardui (Diptera: Tephritidae) in <strong>Canada</strong>. Canadian Entomologist<br />

114.349-357.<br />

Peschken. D.P.; Hunter. J.H.; Thomas. A.O. (1980) Damage in dollars caused by <strong>Canada</strong> thistle in wheat in Saskatchewan. Proceedings <strong>of</strong><br />

the <strong>Canada</strong> Thistle Symposium, Regina, p. 37-43.<br />

Pesehken, D.P.; Johnson, O.R. (1979) Host specificity and suitability <strong>of</strong> Lema cyanella (Coleoptera: Chrysomelida). a candidate for the<br />

biological control <strong>of</strong> <strong>Canada</strong> thistle (Cirsium arvense). Canadian Entomologist III. 1059-1068.<br />

Peschken. D.P.; Wilkinson. A.T.S. (1981) Biocontrol <strong>of</strong> <strong>Canada</strong> thistle (Cinium arvense): releases and effectiveness <strong>of</strong> Ceutorhynchus<br />

litura (Coleoptera: Curculionidae) in <strong>Canada</strong>. Canadian Entomologist 113, 777-785.<br />

Schaber. B.C.; Balsbaugh. E.U .• Jr.; Kantack, B.H. (1975) Releases <strong>of</strong> Altica carduorum (Col.: Chrysomciidae) on <strong>Canada</strong> thistle (Cirsium<br />

arvense) in South Dakota. Entomophaga 20, 325-335.<br />

Story. J.M. (1980) Biological control <strong>of</strong> <strong>Canada</strong> thistle in Montana. Proceedings <strong>of</strong> the <strong>Canada</strong> Thistle Symposium. Regina. p. 128-129.<br />

Thomas. A.O. (1976) Weed survey <strong>of</strong> cultivated land in Saskatchewan. Weed survey series, Publication 76-1, 93 pp.<br />

Thomas. A.O. (198O) Relative abundance <strong>of</strong> <strong>Canada</strong> thistle on cultivated land in <strong>Canada</strong>. Proceedings <strong>of</strong> the <strong>Canada</strong> Thistle Symposium,<br />

Regina. p. 167-181.<br />

Ward, R.H.; Pienkowski, R.L. (1978o) Biology <strong>of</strong> Cassida rubiginosa, a thistle-feeding shield beetle. Annals <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong><br />

<strong>of</strong> America 71(4). 585-591.<br />

Ward. R.H.; Pienkowski, R.L. (1978b) Mortality and parasitism <strong>of</strong> Cassida rubiginosa, a thistle-feeding shield beetle accidentally<br />

introduced into North America. Environmental Entomology 7(4),536-540.<br />

Wilson, R.O. (1981) Effect <strong>of</strong> <strong>Canada</strong> thistle residue on growth <strong>of</strong> some crops. Weed Science 29(2), 159-164.<br />

Winiarska, W. (1973) Observations on the biology <strong>of</strong> Lema cyanella L. Chrysomelidae, Criocerinae. Polskie Pismo Entomologiczne 43(2).<br />

373-382 (in Polish).<br />

*Zetterstedt, J.W. (1847) Diptera Scandinavia. V. VI. Lund. p. 2163-2580.<br />

Zwolfer, H. (1967) Observations on Urophora cardui L. (Trypctidae). Weed projects for <strong>Canada</strong>. Commonwealth Institute <strong>of</strong> Biological<br />

Control Progress Report 14. Delemont. 11 pp.<br />

Zwolfer, H. (1969) Experimental feeding ranges <strong>of</strong> species <strong>of</strong> Chrysomelidae (Col.) associated with Cynareae (Compositae) in Europe.<br />

Commonwealth Institute <strong>of</strong> Biological Conrrol Technical Bulletin 12. 115-130.<br />

Zwolfer, H.; Pattullo. W. (1970) Zur Lebensweise and Wirtsbindung des Distel-Blattkiifers Lema cyanella L. (puncticollis Curt.) (Col.<br />

Chrysomelidae). Anzeiger fur Schiidlingskunde. Pj1anzenschutz, UmweltschUlz. 43(4). 53-59<br />

* The article has not been seen by the author.


Pest Status<br />

Background<br />

Releases and Recoveries<br />

Chapter 33<br />

Urophora sty/ala<br />

F. (Diptera:<br />

Tephritidae) (a) Ecology<br />

Cirsium vulgare (Savi) Ten., Bull Thistle<br />

(Compositae)<br />

P. HARRIS and A.T.S. WILKINSON<br />

Cirsium vulgare (Savi) Ten. is a biennial <strong>of</strong> Eurasian origin that propagates itself only by<br />

seed. It is found in all regions <strong>of</strong> southern <strong>Canada</strong> but is most abundant in Quebec,<br />

Ontario, and British Columbia (Moore & Frankton 1974) on recently disturbed uncultivated<br />

ground or overgrazed permanent pasture. In southern British Columbia, it occurs<br />

along roads, ditches, dykes and fences, and on pastures or range where the sward has<br />

been broken; it is <strong>of</strong>ten troublesome in forage crops in the second year after planting.<br />

The thistle does not withstand cultivation so it is not a pest in arable crops. It may have<br />

one to several flowering stems, usually 0.5 to 2 m high but sometimes 3 m. Plants can be<br />

up to one meter in diameter with as many as 343 seed heads. The sharp prickles on the<br />

leaves discourage livestock from grazing it.<br />

Bull thistle can be controlled with 2,4-D or MCPA at 1 kg/ha; mowing is also effective<br />

if it is done when the plant is reaching maturity; but if cut too early the plant regrows and<br />

sets seed before frost. In spite <strong>of</strong> the ease <strong>of</strong> control, bull thistle persists as a common<br />

weed although it is a minor problem compared with the other thistles targeted for biological<br />

control in this volume.<br />

Surveys <strong>of</strong> the European insect fauna on C. vulgare were made by Zw6lfer (1965) as part<br />

<strong>of</strong> a broad study <strong>of</strong> thistle insects that might be used for biological control in <strong>Canada</strong>. In<br />

North America, the flower buds <strong>of</strong> C. vulgare are attacked by the artichoke plume moth,<br />

Platyptilia carduidactyla (Riley) (Pterophoridae), and the vegetative parts by Entylia<br />

carinata (Forst.) (Membracidae) as well as by a number <strong>of</strong> more polyphagous insects.<br />

There was an absence <strong>of</strong> specialized insects such as Urophora stylata F. (Tephritidae) in<br />

the seed heads. As a result <strong>of</strong> studies by Zw6lfer (1965), Persson (1963), and Redfern<br />

(1968), it was apparent that U. stylata was a promising biological control agent. Thus its<br />

screening and introduction to <strong>Canada</strong> was partly opportunistic. The lack <strong>of</strong> screening<br />

and release <strong>of</strong> other agents for the control <strong>of</strong> C. vulgare reflects the relatively minor pest<br />

status <strong>of</strong> the thistle.<br />

The biology <strong>of</strong> U. stylata is reported by Redfern (1968) and ZwOlfer (1972). The mature<br />

larvae overwinter in the seed heads <strong>of</strong> C. vulgare and pupate in the spring. The univoltine<br />

adults emerge in early summer (June) and the males establish territories on bolting thistle<br />

plants. Shortly before the thistles start to bloom the female visits the plants, mates,<br />

oviposits into the flower buds at a precise stage <strong>of</strong> maturity, and then leaves for another<br />

147


148 P. Harris. and A. T. S. Wilkinson<br />

Table 34<br />

thistle. The eggs are laid in small groups through the bracts at the top <strong>of</strong> the bud. A<br />

second-instar larva hatches in 5-8 days and bores down a corolla tube into the ovariole.<br />

The ovariole and the adjacent receptacle become enlarged and the larva feeds inside this<br />

gall. Lignified tissues which develop on the outside <strong>of</strong> the gall gradually coalesce with<br />

those <strong>of</strong> adjacent galls so that the larvae are eventually enclosed in a hard multilocular<br />

gall.<br />

Zwolfer (1972) and Redfern (1968) reported that approximately one quarter <strong>of</strong> the C.<br />

vulgare heads collected in their studies in western and central Europe were attacked by U.<br />

stylala. This is much lower than the population reached at the best site in British<br />

Columbia. On the European continent, 37% <strong>of</strong> the larvae were parasitized (Zw6lfer<br />

1972) and 24% in Britain (Redfern 1968). No larval or pupal parasitoids were encountered in<br />

<strong>Canada</strong>. In certain regions <strong>of</strong> Europe, strains <strong>of</strong> U. stylala are specialized on certain other<br />

thistles, but no attack <strong>of</strong> other thistles by the strain released has been found in <strong>Canada</strong>.<br />

(b) Releases<br />

The releases <strong>of</strong> U. stylala are summarized in Table 34. The fly established itself readily<br />

at sites with a dense stand <strong>of</strong> C. vulgare. It has subsequently declined or disappeared<br />

where the thistle became sparse following the reestablishment <strong>of</strong> grasses and other<br />

herbaceous perennials as has happened at Danville and lie Jesus, Quebec, and Cranbrook,<br />

British Columbia. At aka, Quebec, the thistle was mowed one year after release<br />

and the colony disappeared. The release sites at Nanaimo and at Cloverdale, British<br />

Columbia, were also destroyed but the fly can be found in the surrounding area. The<br />

current status <strong>of</strong> the other release sites is unknown. The results <strong>of</strong> establishment suggest<br />

that fly dispersal is slow and they require a relatively stable and dense stand <strong>of</strong> C.<br />

vulgare for survival.<br />

Open releases and recoveries<br />

(Savi) Ten. in <strong>Canada</strong>.<br />

<strong>of</strong> Urophora stylala F. on Cirsium vulgare<br />

Year Location Origin No. released Year <strong>of</strong> recovery<br />

1973 Nanaimo, Germany - 174 adults 1974 (site cleared<br />

British Columbia Switzerland and mowed in 1975)<br />

1973 Westham Isle. Germany - 462 adults 1974<br />

British Columbia Switzerland<br />

1973 Cloverdale, Germany - 132 adults 1974<br />

British Columbia Switzerland<br />

1973 Cranbrook, Germany - 459 adults 1974-1976 (no<br />

British Columbia Switzerland thistles or U.<br />

stylala in 1981)<br />

1976 aka, Quebec France - 132 adults Site mowed - no<br />

Austria establishment<br />

1977 Danville, Quebec British Columbia 157 adults 1978<br />

1977 lIe Jesus, Quebec British Columbia 248 adults 1978<br />

1978 New Denver, British Columbia 957 pupae 1979<br />

British Columbia<br />

1978 Williams Lake, British Columbia 2000 pupae 1979<br />

British Columbia<br />

(lots 6069, 7017, 69)


Table 35<br />

Cirs;llm I'ulg(lre (Savi) Ten. , 149<br />

The establishment and build-up <strong>of</strong> U. stylata populations were followed in more detail<br />

at Cranbrook, Nanaimo. and particularly Cloverdale and Ladner. British Columbia<br />

(Table 35). The Cranbrook site was along a power right-<strong>of</strong>-way through forest; Cloverdale<br />

was a permanent pasture on black muck soil in the Fraser delta on which the thistle<br />

grew where the cattle hooves broke the sward. The site was destroyed for development<br />

in 1980, but the fly survives in the area. At Ladner, also in the Fraser delta. the thistle<br />

was confined to the edge <strong>of</strong> a drainage ditch through an intensely cultivated area used<br />

largely to grow seed sugar beet and carrots. The populations were followed by sampling,<br />

at the end <strong>of</strong> the summer, all the seed heads on thistle plants growing at approximately 5<br />

m intervals in 4 directions from the release point. The exception was the linear stand at<br />

Ladner. Samples were taken to the edge <strong>of</strong> the colony or the site. whichever was least.<br />

Size and spread <strong>of</strong> U. stylata F. colonies at release sites in British Columbia.<br />

Heads/plant No. heads % heads No. larvaeJgall Colony Distance (m) to near-<br />

Location Year ± S.E.M. sampled galled ± S.E.M. radius m est thistle ± SD<br />

Cranbook 1973 21.2 ± 6.0 525 6.7 20 II.R ± 11.1<br />

197-1 12.1 ± 1.8 320 2.8 2.1 ± 0.4 100 2.9 ± 2.9<br />

1976 6.2 ± 1.11 19-1 3.6 2.0 ± 0.4 80 thistle scarce<br />

19111 No thistles prescnt and no U. srylata found on nearby roadside C. vulgart'<br />

Nanaimo 1973 9.4 ± 2.1 113 3.5 30 0.7 ± 0.8<br />

1974 8.7 ± 1.5 180 42.8 3.0 ± 0.3 100 0.2 ±0.3<br />

1975 Area bulldoled, sown to grass and surrounding thistle cut<br />

Cloverdale 1973 53 30.2 3.0 ± 1.1<br />

1974 29.3 ± 3.5 299 41.5 3.2 ± 0.2<br />

1975 32.2 ± 5.0 10M 15.9 140 2.4 ± 3.4<br />

(299 plantslha)<br />

1976 39.7 ± 9.7 1110 65.3 140 + 36.1 ±5.2<br />

1977 39.8 ± 5.9 1\53 84.7 9.1 ± 0.3 200 + 3.5 ± 4.1<br />

1978 38.4 ± 5.7 1283 95.8 13.7 ± 0.6 250+ 2.4 ± 0.4<br />

1979 34.7 ± 7.2 520 92.7<br />

1980 Site destroyed<br />

(site 2) 1978 511.0 ± 19.4 5811 92.2<br />

1979 60.1 ± 14.5 402 83.6<br />

Ladner 1973 22.5 ± 4.7 -149 12.2 3.1 ± 0.3 100 0.3 ± 0.3<br />

1974 Not sampled - all but 2 thistles mowed<br />

1975 54.1 ± 9.2 1165 24.4 125 1.0 ± 1.1<br />

1976 38.6 ± 6.9 1313 5.7 3!!6+ 2.8 ± 6.3<br />

1977 80.6 ± 15.7 2338 6.0 160 + 1.7 ± 1.7<br />

19711 59.7 ± 10.6 1851 31.7 ISO + 11.9 ± 1.3<br />

1979 61.2 ± 10.9 2348 46.3 14.2 ± 5.2<br />

1980 41.8 ± 5.9 1918 39.9<br />

1981 55.3 ± 11.8 775 26.3<br />

1982 39.6 ± 9.0 912 38.8<br />

The results (Table 35) show that at Cloverdale the fly population increased to attack<br />

over 90% <strong>of</strong> the heads formed. The number <strong>of</strong> larvae per head increased with the percentage<br />

<strong>of</strong> heads attacked (Equation 8. Table 36). This is also apparent from Table 35: at<br />

low densities <strong>of</strong> the fly. a density <strong>of</strong> 2-3 larvae per gall was found. but at high fly<br />

densities the average increased to around 14 larvae per gall with maximum numbers <strong>of</strong><br />

over 30 larvae. Presumably the high larval numbers are the result <strong>of</strong> multiple ovipositon.<br />

The data for Ladner were excluded from this analysis where it is suspected that the<br />

inability <strong>of</strong> the fly to attack more than half the heads resulted from their loss over the<br />

surrounding fields which were treated with insecticide during the season. Thus although


150 P. Harris and A. T. S. Wilkinson<br />

Table 36<br />

Evaluation <strong>of</strong> Control Attempts<br />

ElTects <strong>of</strong> U. stylata<br />

on the number <strong>of</strong><br />

seed heads on<br />

C. vulgare plants<br />

the high number <strong>of</strong> larvae/gall in 1979 indicated a high density <strong>of</strong> flies in early summer.<br />

the percentage <strong>of</strong> heads attacked suggests that those forming late in the season escaped<br />

the fly.<br />

The average distance <strong>of</strong> the sampled thistle to its neighbour (Table 35) gives an idea <strong>of</strong><br />

density although. except in one instance. this cannot be related to the number <strong>of</strong> thistles<br />

per unit area as the distribution contagion was not measured. No effect <strong>of</strong> the fly on the<br />

distance between thistles was apparent; however as C. vulgare seed was found to be still<br />

viable after approximately 36 years under forest cover (Harrington 1972). it will take a<br />

number <strong>of</strong> years <strong>of</strong> reduced input into the soil seed-bank before the density <strong>of</strong> the thistle<br />

is affected.<br />

Evaluation <strong>of</strong> the effect <strong>of</strong> U. stylata F. on C. vulgare (Savi) Ten.<br />

Regression<br />

1. No. heads on unattacked plants 0.40 x + 2.0<br />

2. No. heads on attacked plants 0.32 x + 6.2<br />

3. No. plump seedlhead from unattacked plants 3.66 r - 18.7<br />

4. No. plump seedlhead from attacked plants -0.12 r + 67.7<br />

5. Mean no. larvaelhead in head size classes 0.16 r - 1.2<br />

6. Maximum number <strong>of</strong> larvaelhead in head<br />

size cI asses 0.24 r + 2.8<br />

7. Live larval weight in mg -0.08 n + 9.5<br />

8. No. larvaclgall 0.11 z + 0.3<br />

9. No. larvaelhead -0.006 x + 9.6<br />

10. Total seed production from unattacked plants 55.51 x + 235.2<br />

11. Total seed production from attacked plants 17.69 x + 571.4<br />

Where x = dry weight <strong>of</strong> plant in g<br />

r = radius <strong>of</strong> receptacle attachment to the involucre in nun<br />

z = the percentage <strong>of</strong> heads attacked<br />

n = no. larvae/head<br />

Cor.Coef. n P<br />

0.97 25


The number <strong>of</strong> seeds<br />

per head vs. head<br />

size and the number<br />

<strong>of</strong> U. sty/ala<br />

larvae present (a) Methods<br />

Cirsium vulgare (Savi) Ten.. 151<br />

recorded for each plant, as well as the number <strong>of</strong> heads attacked, and the number <strong>of</strong> U.<br />

stylolo larvae per gall. Attack by U. stylala enlarges the receptacle but does not affect<br />

the base where it is attached to the involucre, so its diameter was used to compare the<br />

size <strong>of</strong> the productive heads on the attacked and un attacked plants.<br />

(b) Results<br />

An average <strong>of</strong> 88% <strong>of</strong> the productive heads was galled on the attacked plants. This<br />

percentage is higher than that obtained from the transect samples (Table 35), but as a<br />

range <strong>of</strong> thistle sizes was selected from near the center <strong>of</strong> the colony, it was not a random<br />

sample. The 564 galled heads contained an average <strong>of</strong> 10.3 ± 0.3 SEM larvae. This is also<br />

slightly higher than the mean in Table 35.<br />

Plant weight accounted for most <strong>of</strong> the variation in the number <strong>of</strong> seed heads on both<br />

the attacked and unattacked plants (Equations 1 and 2. Table 36). The level <strong>of</strong> attack on<br />

individual plants ranged from an average <strong>of</strong> 1. 7 larvae/head to 18.1 larvae/head, but the<br />

attack level was not affected by plant size (Equation 9, Table 36). The slopes <strong>of</strong> the<br />

regression lines (Equations 1 and 2) suggest that attack reduced the number <strong>of</strong> heads:<br />

plants <strong>of</strong> 100 g had 9.5% fewer and those <strong>of</strong> 200 g had 14.6% fewer than the unattacked<br />

plants. However, the slopes were not significantly different and the differences in the<br />

averages were partly balanced by a slightly larger head size on the attacked plants<br />

(average diameter 13.7 ± 0.07 mm compared with 12.8 ± 0.06 for the unattacked plants).<br />

It is concluded that the receptacle area in the two groups <strong>of</strong> plants was practically identical.<br />

This was in contrast to the results <strong>of</strong> the seed-head Urophora spp. on diffuse knapweed<br />

(Harris 1980) where the number <strong>of</strong> heads and plant size were reduced by attack.<br />

Heads from both attacked and unattacked plants were bagged shortly before the pappus<br />

was ready to fly. Seed head size was measured as the radius <strong>of</strong> the receptacle attachment<br />

to the involucre as previously described, and the plump seed. separated with a blower,<br />

was counted for each head. Plump seed rather than germinated seed was used for<br />

regression analysis: both produced similar trends but the correlations were higher with<br />

the former. Part <strong>of</strong> the problem was a high variability <strong>of</strong> germination between heads.<br />

Anderson (1968) reported 40% germination at 20°C although this was increased to 80%<br />

with temperatures alternating between 20-30°C. We obtained 66.2 ± 2.0% germination<br />

<strong>of</strong> plump seed from unattacked plants at room temperature and 52.4 ± 4.8% from<br />

attacked plants.<br />

(b) Results<br />

On unattacked plants, large heads produced more plump seed than did small ones<br />

(Equation 3, Table 36) as is normal in composites. In contrast on the attacked plants,<br />

seed production decreased with head size (Equation 4, Table 36) although the slope and<br />

the correlation coefficient were so low that the production <strong>of</strong> 62 seeds for an average<br />

head <strong>of</strong> 13.7 mm diameter can be used as a constant for all heads. This means that for the<br />

plant it is now more productive to develop several small heads rather than a few large<br />

ones.


152 P. Harris and A. T. S. Wilkinson<br />

Calculation <strong>of</strong> seed<br />

production by<br />

attacked and<br />

unattacked plants<br />

Recommendations<br />

Literature Cited<br />

The effect <strong>of</strong> head size on seed production in the attacked plants is nullified because<br />

the number <strong>of</strong> U. sty/ala also increases with head size (Equation 5, Table 36). The mean<br />

number <strong>of</strong> larvae per head was considerably lower than the maximum number in the<br />

heads <strong>of</strong> each size (Equation 6, Table 36). Thus for a head with a diameter <strong>of</strong> 14 mm the<br />

average number from the regression was 6.6 larvae while the maximum was 14.6 larvae.<br />

Thus the larval population is well below the physical limits <strong>of</strong> the thistle heads. There<br />

was a rather small effect <strong>of</strong> crowding on larval weight. A regression <strong>of</strong> live weight <strong>of</strong> 980<br />

larvae from heads with 1-33 larvae in a gall showed that one larva per gall averaged 9.4<br />

mg compared with 6.7 mg for those in heads with 33 larvae. The slope <strong>of</strong> the regression is<br />

highly significant, although as indicated by the small correlation coefficient there was a<br />

large variation in larval weights at all densities (Equation 7, Table 36). For the purpose <strong>of</strong><br />

the calculations in this paper the size and hence the effect <strong>of</strong> a larva was assumed to be<br />

the same whether it was single or crowded.<br />

(a) Methods<br />

Seed production for each head on the unattacked plants was determined from regression<br />

Equation 3 (Table 36). These were summed and the total for each plant regressed on<br />

plant weight so that average production could be determined for a plant <strong>of</strong> any size. The<br />

same was done for the attacked plants except that the heads were assumed to produce 62<br />

seeds regardless <strong>of</strong> size.<br />

(b) Results<br />

The production <strong>of</strong> 4398 seeds for a plant <strong>of</strong> 75 g (Equation 10, Table 36), is close to the<br />

figure <strong>of</strong> 4000 seeds/plant given by Salisbury (1964) for the thistle in Britain. The attacked<br />

plant <strong>of</strong> this size produced 1898 seeds (Equation 11, Table 36), a reduction <strong>of</strong><br />

57%. For a plant <strong>of</strong> 150 g, the reduction was 62%.<br />

The calculations were made for a U. Sly/ala population infesting 88% <strong>of</strong> the thistle<br />

heads. This is slightly below the population plateau at Cloverdale, so the reduction in<br />

plump seed production at this site should be at least 60%. If the difference in the germination<br />

between attacked and unattacked plants is accepted, the reduction is over 65%.<br />

Under field conditions germination and seedling survival seem to depend on breaks in<br />

the sward rather than competition within the C. vulgare population. This means that the<br />

reduction in thistle density should be similar to the reduction <strong>of</strong> seed as soon as the soil<br />

seed-bank has reached its new equilibrium.<br />

(1) U. stylala should be distributed to sites across <strong>Canada</strong> with stable populations <strong>of</strong><br />

C. vulgare; but there is little point in releasing it against temporary outbreaks.<br />

(2) The introduction <strong>of</strong> additional biological control agents does not appear to be<br />

warranted as most <strong>of</strong> the problems with C. vulgare are either temporary following the<br />

clearing <strong>of</strong> land or the result <strong>of</strong> poor farming practice.<br />

Anderson. R.N. (1968) Germination and establishment <strong>of</strong> weeds for experimental purposes. Weed Science <strong>Society</strong> <strong>of</strong> America, 636 pp.<br />

Harrington. J.F. (1972) Seed storage and longevity. In: Kozlowski T.T. (Ed.) Seed biology vol. III. Academic Press. pp. 145-245.


CiTSiu11I J'lllgaTe (Savi) Ten. , 153<br />

Harris, P. (19BO) Effects <strong>of</strong> UropllOra affillis Frlld. and U. lfuaelrifruciata (Meig.) (Diptera: Tephritidae) on Centaurea diffusa Lam. and C.<br />

maculosa Lum. (Composilae). 'Zritse/rri[t frlr allgewalldte Emomologie 90, 190-210.<br />

Moore, R.J.; Frankton, C. (1974) TIle thistles <strong>of</strong> Canuda. Agricullllre CU/lelda MOllograplr 10, 111 pp.<br />

Persson, P.I. (1963) Studies on the biology and huvill morphology <strong>of</strong> some trypetids (Dipt.). Opuscula Entomologica 28, 3-69.<br />

Redfern, M. (1968) The natural history <strong>of</strong> spear thistle hends. Field SlIldies 2.669-717.<br />

Salisbury. E. (1964) Weeds lind aliens. Collins, 384 pp.<br />

Zwolfer, H. (1965) A prcliminary list <strong>of</strong> phytoplurgous insects IIttllcking wild Cynareae species in Europe. Commonwealtll Instillete <strong>of</strong><br />

Biellogical Control Tee/mimi B,llielill 6, 81-154.<br />

Zwolfer. H. (1972) Invcstiglltions on Ur(lplwra sty/ala Fabr •• a possible agent for Ihe biological control <strong>of</strong> Cinium vIClgare in <strong>Canada</strong>.<br />

Weed projects for C:lDmJa. Commollwcallir /tlStilllle <strong>of</strong> Biological Control Progress Report 29. 20 pp.


Blank Page<br />

154


Pest Status<br />

Background<br />

Biological Control<br />

Chapter 34 155<br />

Convolvulus arvensis L., Field Bindweed<br />

(Convolvulaceae)<br />

M.G. MAW<br />

Field bindweed. Convolvulus arvensis L.. a deep rooted perennial <strong>of</strong> Eurasian origin. is<br />

a serious problem in many areas and has been classed as the worst weed <strong>of</strong> the prairie<br />

and the Great Plains States (Bakke et al. 1939). It is a serious weed in Ontario and<br />

Quebec. especially in corn. and although it is <strong>of</strong> less importance in the western provinces<br />

some local areas are badly infested.<br />

Losses from bindweed competition vary with the crop and cultural methods. Rye<br />

yields may be reduced by 20% while grain sorghum yields may be reduced by 78% . On<br />

bindweed free land, wheat. barley. and oats yielded respectively 350.576. and 515 kg per<br />

ha more than on infested fields (Wiese & Phillips 1976). Fast growing crops that shade<br />

the ground may compete effectively with the weed. and winter wheat on the drier parts<br />

<strong>of</strong> the Great Plains is a good competitor because winter and spring growth occurs when<br />

bindweed is dormant. Summer crops may however be severely stunted or even killed by<br />

severe bindweed competition (Weise & Phillips 1976).<br />

Field bindweed was first noticed in America in Virginia as early as 1739. and by 1900 it<br />

was well established and recognized as a serious weed in most <strong>of</strong> the western states. In<br />

addition to a progressive spread across America. the weed may have been introduced<br />

several times. For example. it may have been introduced directly into Kansas with<br />

wheat brought from the Ukraine by immigrants about 1875 (Weise & Phillips 1976).<br />

Bindweed is difficult to control because <strong>of</strong> its extensive root system and prolific seed<br />

production. Roots and rhizomes may cover an area up to 6 m in diameter (Frazier 1943)<br />

and extend down 9 m below the soil surface (Phillips 1978). In addition roots store a<br />

range <strong>of</strong> sugars. starches. and proteins that can support the plants over long periods<br />

even though top growth may be repeatedly destroyed (Weise & Phillips 1976). Over 11<br />

million seeds may be produced per hectare and they may remain viable in soil for up to 30<br />

years (Timmons 1949).<br />

Seedlings develop deep taproots in a few weeks and in 9 to 11 weeks spread radially by<br />

lateral roots. Within one season a single plant may form a patch <strong>of</strong> more than 3 m in<br />

diameter (Weise & Phillips 1976).<br />

Control <strong>of</strong> bindweed but not eradication can be achieved through an integrated programme<br />

<strong>of</strong> frequent cultivation and selected herbicides. However, with the growing<br />

concern <strong>of</strong> the overload <strong>of</strong> chemicals in the environment, their cost. the cost <strong>of</strong> energy,<br />

and costs in lost production there has been interest in biological control. Mohyuddin<br />

(1969) listed the insects from Calystegia spp. and Convol"lIll1s spp. found by him in<br />

Ontario and by others in other parts <strong>of</strong> the world. and the University <strong>of</strong> California and<br />

the USDA jointly mounted a project to survey the Mediterranean area for natural<br />

enemies <strong>of</strong> Convolvulus (Rosenthal 1980). In all, 140 species <strong>of</strong> insects. three species <strong>of</strong><br />

mites, and three fungi were found attacking C. arvensis and other closely related Convolvulaceae.<br />

Of these, a mite, Eriophyes sp.; a bruchid. Spermophagus sericelts Ge<strong>of</strong>f.; a


156 M. G. Maw<br />

Discussion<br />

Table 37<br />

Table 38<br />

chrysomelid. Ga/eruca TUfa Germ.; and a mildew. Erysiphe convolvuli DC.. showed the<br />

most potential as control agents (Rosenthal 1980). None. however. was found to be<br />

strictly monophagous.<br />

Little biological control work has been done in <strong>Canada</strong>. However. limited attempts to<br />

augment existing populations or to establish new colonies in Alberta and British<br />

Columbia with insects from Saskatchewan and Ontario (Table 37 and Table 38) were not<br />

successful.<br />

Collection and release in <strong>Canada</strong> <strong>of</strong> chrysomelid insects in control <strong>of</strong> COlU'O/VIl/US arvensis<br />

L. projects<br />

Species Number Year Source Release or Lab Study<br />

Chirida guttala 114 1969 <strong>Canada</strong> British Columbia (study)<br />

(01.) 125 1970 Ontario British Columbia (study)<br />

236 1979 Saskatchewan Alberta (release)<br />

C/lelymorpha cassidea (F.) 97 1979 Saskatchewan Alberta (release)<br />

Melriona bie%r (F.) 18 1969 <strong>Canada</strong> British Columbia (study)<br />

7 1970 Ontario British Columbia (study)<br />

Metriona purpurata Boh. 224 1979 Saskatchewan Alberta (release)<br />

Collection and release in <strong>Canada</strong> <strong>of</strong> chrysomelid insects in control <strong>of</strong> Convo/vulus sepiwll<br />

L. projects<br />

Species<br />

Chirida gllttala (01.)<br />

Melriona bie%r (F.)<br />

Number<br />

120<br />

200<br />

Year<br />

1971<br />

1971<br />

Source<br />

Ontario<br />

Ontario<br />

Release<br />

British Columbia<br />

British Columbia<br />

Bindweed competes with crop plants for nutrients and water, and thus lowers yields. It<br />

twines around plants, smothering them and interfering with harvesting, and is a haven<br />

for crop pests (Rosenthal 1980}. It is considered to be the 12th most important weed in<br />

the world and the 14th most important weed in the United States (Rosenthal 1980). It is a<br />

serious weed in Ontario and Quebec but a lesser problem in the western provinces. It is<br />

mainly a weed in cultivated fields and so not a good candidate for biological control.<br />

A large number <strong>of</strong> insects are associated with bindweed in North America. but only<br />

sporadically in localized areas is appreciable damage done to the plant (Maw unpublished<br />

data). Also, it appears that most insect damage occurs when bindweed is<br />

supported by other plants or fences, rather than in the prostrate growth in fields.<br />

Attempts to move Chirida gllliala, Melriona bicolor and M. purpurala, and Che/ymorpha<br />

cassidea into new areas were unsuccessful. This may indicate that these species<br />

have reached their geographic limits.<br />

Rosenthal & Buckingham (1982) found a number <strong>of</strong> European organisms that may<br />

be useful as control agents <strong>of</strong> bindweed in North America. Of these. the arthropods<br />

Ga/eruca rufa, Spermophagus sericeus, and Eriophyes sp. probably have the most<br />

potential (Rosenthal 1980) and several other species may also be useful (Rosenthal &


Blank Page<br />

158


Pest Status<br />

Table 39<br />

Chapter 35<br />

Euphorbia esula-virgata complex, Leafy<br />

Spurge and E. cyparissias L., Cypress<br />

Spurge (Euphorbiaceae)<br />

P. HARRIS<br />

Leafy and cypress spurge. Euphorbia esu/a'I'irgala complex and E. cypar/ss/as<br />

L.. are herbaceous perennials <strong>of</strong> European origin that have become problem species in<br />

parts <strong>of</strong> North America. Leafy spurge is particularly serious on the Canadian prairies<br />

and the north central United States. while cypress spurge is most prevalent in Ontario<br />

and Quebec. Both spurges are unpalatable to most grazing animals except sheep and<br />

displace other herbaceous vegetation. In two leafy spurge stands investigated in Sas·<br />

katchewan in 1981. there was 65-72% less grass within the stand than just outside it.<br />

Cypress spurge on a thin limestone soil at Braeside. Ontario. comprised 56% <strong>of</strong> the<br />

herbaceous vegetation in a dry year and 26% in a wet year (Table 39). Cypress spurge is<br />

Density <strong>of</strong> Hy/es euphorbiae (L.) larvae, dry weight <strong>of</strong> E. cyparissias L. and total<br />

herbaceous vegetation at Braeside. Ontario.<br />

Year No. H. euphorbiae Dry weight <strong>of</strong> Dry weight <strong>of</strong> all<br />

larvae/m 1 E. cyparissias glml vegetationlm l<br />

1968<br />

6.8 x 10-)<br />

1969 1.75 x 10- 1<br />

56.6 216.0<br />

1970 0.4 44.7 79.2<br />

1971 1.0 73.1 147.1<br />

1972 0.3 91.9 213.1<br />

1974 0.6<br />

1975 2.3<br />

1976 0.4<br />

also <strong>of</strong> concern as extremely small amounts <strong>of</strong> its latex are intensely irritating to the<br />

eyes. The provincial and most municipal governments <strong>of</strong> the Canadian prairies have an<br />

active programme for the control <strong>of</strong> leafy spurge. For example in the early 1950s,<br />

Saskatchewan spent around $100 OOO/year against persistent perennial weeds and by 1979<br />

the amount had risen to $150 OOO/year, with leafy spurge being the main target. The sum<br />

covers the cost <strong>of</strong> treatment on provincial and municipal lands and roadsides but only<br />

the cost <strong>of</strong> the chemical on private lands. Thus the actual cost <strong>of</strong> controlling leafy spurge<br />

was considerably higher. In Manitoba in 1981. $38649 was spent on leafy spurge control<br />

by the provincial government <strong>of</strong> which $32 890 represented the cost <strong>of</strong> the chemical (E.<br />

Johnson. 1982. personal communication). In Saskatchewan in 1982, the full cost was<br />

$640Iha <strong>of</strong> spurge treated as the infestations are scattered or on rough terrain.<br />

In the first twenty years <strong>of</strong> the control programme in Saskatchewan the chemicals<br />

used were soil stcrilants such as sodium chlorate and various borates either alone or in<br />

combination with other herbicides. These were both expensive and ecologically un-<br />

159


160 P. Harris<br />

Background<br />

desirable. In 1971, the treatment <strong>of</strong> 65 ha with borascu (disodium tetraborate and 5<br />

bromo-3-sec-butyl-6-methylurate) cost $15 727 while the treatment <strong>of</strong> 86 ha with<br />

picloram (4-amino-3,5,6-trichlorpiclinic acid) cost $12 073 (SaskatchewllO Production<br />

and Marketing Branch 1972). Not only was picloram cheaper but the grass cover<br />

remaining was greatly preferable to the bare patches following treatment with soil<br />

sterilants. Thus in the past ten years picloram has been used almost exclusively on<br />

uncultivated stands except in Manitoba where its use is more restricted.<br />

The effectiveness <strong>of</strong> the chemical control programme is not impressive. Eleven<br />

Saskatchewan municipalities were resurveyed for leafy spurge llfter approximately 30<br />

years <strong>of</strong> chemical control. The weed has disappeared on cereal land with normal cultivation<br />

and application <strong>of</strong> 2.4-D, while on uncultivated land the area infested with spurge has<br />

more than doubled. These two trends almost cllOcelled eaeh other in the municipalities<br />

sampled.<br />

In Manitoba. where little picloram has been used, the increase has been large: from<br />

3236 ha in 1952 (Craik 1953) to an area estimated by Johnson (1982, personal communication)<br />

to be 46 693 ha in 1981. The difference in spurge increase between Saskatchewan<br />

and Manitoba is attributed entirely to the use <strong>of</strong> picloram; but against the<br />

benefits <strong>of</strong> the herbicide is the hazard that large scale continued use will contaminate the<br />

ground run<strong>of</strong>f water since the herbicide is subject to little microbial breakdown.<br />

The largest cypress spurge infestation in Ontario is on thin uncultivated limestone<br />

soils at Braeside. The species appears to have been introduced in 1870 and by 1981 had<br />

spread, initially with highway construction. to about 14000 ha on two limestone ridges.<br />

The main cause for concern is that the present infestation is only 15 km away from the<br />

360 000 ha Smith Falls Limestone Plain.<br />

The discussion <strong>of</strong> the biological control <strong>of</strong> leafy spurge by Harris et al. (in press) lists<br />

many insects and pathogens that are restricted to the genus Euphorbia. The difficulty is<br />

finding species that will survive on North American leafy spurge which seems to be a<br />

species complex. Specialized E. esula insects such as the aphid Acyrthosipholl neerlandicum<br />

HRL and the clear-winged moth. Chamaesphecia tentilrediniformis. do not survive on the<br />

common North American leafy spurge. To find insects that will accept the spurge. it has<br />

been necessary to seek those that can survive on several Euphorbia spp. Conversely.<br />

there are several spurges that should not be damaged in North America: E. antisyphilitica<br />

(a source <strong>of</strong> high quality wax in Mexico), E. pu/cherrima (the ornamental poinsettia). E.<br />

lathyris (a plant suggested as a source <strong>of</strong> hydrocarbon (Calvin 1978» and native North<br />

American spurges in general. Hence the biological control <strong>of</strong> leafy spurge is not without<br />

its difficulties even though the presence <strong>of</strong> large dense stands that arc almost unutilized by<br />

phytophagous organisms suggests that it is an extremely amenable target for biological<br />

control.<br />

About three quarters <strong>of</strong> the insects attacking Euphorbia in Europe are restricted to the<br />

genus. This is a higher proportion than for most other herbaceous plants. For example. at<br />

least three quarters <strong>of</strong> the insects that feed on the genus Solidago attack other plant<br />

genera (Harris et al .• in press). It appears that the degree <strong>of</strong> specialization in the insect<br />

guild on spurge is determined by the presence <strong>of</strong> enzymes and toxins that make it difficult<br />

for utilization by unspecialized insects and this extends to individual spurge species. For<br />

example. Ellis & Lennox (1941) reported the proteolytic enzyme euphorbain from E.<br />

lathyris but that E. peplus contained little or none <strong>of</strong> it. That some <strong>of</strong> the compounds are<br />

extremely toxic has been long known. For example the Geoponika (6th or 7th century)<br />

records that tithymalus (possibly E. cyparissias) mixed in a bait could be used to control<br />

mice by making them blind. I can confirm that minute amounts <strong>of</strong> E. cyparissias latex in


Releases and Recoveries<br />

HyJes euphorbiae<br />

(L.) (Lepidoptera:<br />

Sphinguidae) (a) Ecology<br />

Euphorbia esula-I'irgata complex. 161<br />

the eye are extremely painful. In the 12th century. Ibn Al Awan (Clemont-Mullet 1864)<br />

described the use <strong>of</strong> a spurge extract for repelling and killing insects. Portier & Busnel<br />

(1951) found that E. cyparissias latex contained a substance that produced violent muscle<br />

tetany in most insects although the specialized spurge hawkmoth larvae. Hyles euphorbiae<br />

(L.). recovered without ill effects. Geilser (1955) reported that the effects <strong>of</strong> the<br />

substance were similar to those <strong>of</strong> acetylcholine in that atropine was antagonistic to it.<br />

Such differences presumably account for the specialization <strong>of</strong> some <strong>of</strong> the Chamaesphecia<br />

spp. moths to a single Euphorbia sp. For biological control purposes. this makes the<br />

precise identification <strong>of</strong> the target leafy spurges <strong>of</strong> North America an urgent matter. As<br />

this has so far proved difficult on morphological grounds perhaps a chemical or isozyme<br />

classification should be attempted.<br />

H. euphorbiae. the spurge hawkmoth. is a large defoliator <strong>of</strong> Euphorbia species in the<br />

subgenus esula in south and central Europe. northern India. and central Asia. The pupae<br />

can survive cold winters (Harris & Alex 1971) and its northern distribution seems to be<br />

limited by a requirement for high summer temperatures as the larvae feed little and<br />

eventually die if kept below 14°C. Also for unknown reasons the moth is uncommon south<br />

<strong>of</strong> the Mediterranean. H. euphorbiae has three generations a year in the south <strong>of</strong> its range<br />

and one in the north.<br />

The moth was selected a a biological control agent as it appeared to have the capability<br />

<strong>of</strong> defoliating spurge stands over large areas. Indeed it was a major pest <strong>of</strong> commercially<br />

grown E.lathyrisin the USSR (Malyuta 1934). The moth was approved for release in 1965<br />

and subsequently attempts were made to establish it in both <strong>Canada</strong> and the United<br />

States.<br />

(b) Releases<br />

Large numbers <strong>of</strong> H. euplrorbiae were bred in the laboratory and released. mostly as<br />

larvae. in spurge stands across <strong>Canada</strong> (Table 40). The stock used was from scattered<br />

collecting in Switzerland. France, and Germany on E. cyparissias and E. seguierana.<br />

Other larvae presumably from these hosts were purchased from collectors in East Germany.<br />

Initial establishment was disappointing and the first evidence <strong>of</strong> breeding was 68 widely<br />

scattered nearly mature larvae recovered at one <strong>of</strong> four release sites at Braeside, Ontario<br />

in 1968. These were collected and two generations bred from them over winter in the<br />

laboratory with the result that over 5000 <strong>of</strong> their progeny were released at Braeside in the<br />

spring <strong>of</strong> 1969. The only other site where a few field bred larvae were collected was at<br />

Sidney. Ontario, but this colony has not been visited since 1972.<br />

Most <strong>of</strong> the releases <strong>of</strong> H. euplrorbiae larvae made in 1966 and 1967 suffered a<br />

mortality <strong>of</strong> over 95% during the subsequent two weeks. The exception was a series <strong>of</strong><br />

releases made in September at Sidney. Ontario. and one release at Braeside. Ontario.<br />

where 23% <strong>of</strong> the larvae survived this period. The development <strong>of</strong> the larvae released<br />

late in the season was so slow that many <strong>of</strong> them failed to pupate and died. Forwood<br />

(1977) with a release <strong>of</strong> 51 larvae on leafy spurge in Nebraska found that they had all<br />

disappeared within seven days. The reason for the larval mortality was predation by


162 P. Harris<br />

Table 40<br />

Table 41<br />

Open releases and recoveries <strong>of</strong> Hyles euphorbiae (L.) against Euphorbia spp. in <strong>Canada</strong>.<br />

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

Year Euphorbia sp. Location No. and Stage Recovery<br />

1965 Cypress spurge Sidney. Ontario 206 larvae<br />

1966 Leafy spurge Jameson, Saskatchewan 1200 larvae<br />

Wilmer, British Columbia 300 larvae<br />

Miami. Manitoba 195 larvae<br />

14 pupae<br />

Hartney. Manitoba 186 larvae<br />

Cypress spurge Sidney. Ontario 3606 larvae<br />

Jones Falls. Ontario 550 larvae<br />

1967 Leafy spurge Wilmer. British Columbia 1900 larvae<br />

Kamloops, British Columbia 3000 larvae<br />

Moose Jaw Creek. Saskatchewan 979 larvae<br />

Balgonie, Saskatchewan 361 larvae<br />

Nr. Balgonie, Saskatchewan 3000 larvae<br />

Milestone, Saskatchewan 901 larvae<br />

Miami, Manitoba 1100 larvae<br />

Picton, Ontario 825 larvae<br />

1967 Cypress spurge Braeside, Ontario 13680 larvae 1968<br />

1968 Cypress spurge Sidney, Ontario 1565 larvae<br />

Braeside, Ontario 2801 larvae<br />

1969 Cypress spurge Braeside, Ontario 5210 larvae<br />

1973 Cypress spurge Hornings Mills, Ontario 600 larvae 1982<br />

1973 Leafy spurge Guelph, Ontario 200 larvae no recovery<br />

1974 Leafy spurge Regina Beach, Saskatchewan 37 larvae<br />

600 pupae<br />

1978 Leafy spurge Cardston, Alberta 229 larvae<br />

ants, mice, carabids, wasps, pentatomids, and spiders (Harris & Alex 1971). At three<br />

sites in Ontario the percentage survival <strong>of</strong> larvae after four days decreased almost<br />

directly with the average number <strong>of</strong> ants caught in a standard series <strong>of</strong> pitfall traps<br />

during the period (Table 41).<br />

The 68 larvae found in 1968 and the 179 found in 1979 that had not been released that<br />

summer were expressed on a m l basis as if all the larvae were confined to a hectare. In<br />

most <strong>of</strong> the subsequent years the survey was restricted to 100 m l samples taken in early<br />

August at paced intervals from a hectare within the release field but in 1975 the samples<br />

were taken from the adjacent roadside which usually supported a higher density <strong>of</strong><br />

larvae.<br />

Relationship between the number <strong>of</strong> ants in pitfall traps and the survival <strong>of</strong> H. euphorbiae<br />

(L.) larvae in a four day period<br />

Place<br />

Braeside<br />

Sidney<br />

Picton<br />

No. ants/trap<br />

5.4<br />

35.9<br />

58.7<br />

% H. euphorbiae<br />

larvae recovered<br />

34.5<br />

20.5<br />

11.8


166 P. Harris<br />

Table 43<br />

in East Austria with 3-5 larvae per plant. Some field mortality was noted: at one site on<br />

E. cyparissias, 31 % <strong>of</strong>the eggs failed to hatch although they developed embryos and 9%<br />

<strong>of</strong> the larvae died in the early instars. The population in part seems to be controlled by<br />

oviposition behaviour as multiple oviposition into a stem is rare and when it does occur<br />

only one larva survives to enter the root. In choice tests O. erythrocephala attacked<br />

Canadian leafy spurge more <strong>of</strong>ten than E. esula, but there was no indication that<br />

association with a particular spurge species in the field had formed a host race.<br />

(b) Releases<br />

Authority to release O. erythroceplwla was obtained in October 1979. The stock was<br />

field collected as beetles in Switzerland on E. segllierana and E. cyparissias, and in<br />

Austria on E. cyparissias and E. eSllla (Table 43). The sexes were separated on<br />

collection and then released together at the release site in <strong>Canada</strong> later.<br />

Open releases and recoveries <strong>of</strong> Oberea erythrocephala (Schr.) on E. esula-virgala complex<br />

in <strong>Canada</strong>.<br />

Release Date No. and Stage Source Release Site Recovery<br />

15 Oct 79 30 larvae Switzerland Caronport, Saskatchewan 1 <strong>of</strong> 12 alive<br />

May 80<br />

27 Jun 80 70 males Switzerland Caronport, Saskatchewan 15 larvae Sep 80<br />

60 females 31 larvae Sep 81<br />

3 Jul80 70 males Switzerland Cardston. Alberta 17 larvae Sep 80<br />

53 females o larvae Aug 81<br />

14 Jul 80 49 males Switzerland Jameson, Saskatchewan 1 larva Sep 80<br />

57 females o larvae Aug 81<br />

26 Jun 81 55 males Switzerland Weyburn, Saskatchewan 14 larvae Aug 81<br />

(8 on plants<br />

attacked in lab)<br />

30 Jun 81 59 males Switzerland Lauder, Manitoba 12 larvae Aug 81<br />

36 females<br />

10 Jul 81 90 adults Switzerland Manitou Lake, Saskatchewan 13 larvae Aug 81<br />

& Austria<br />

5 Aug 81 17 larvae Austria Caronport, Saskatchewan<br />

In 1979,30 leafy spurge plants that had been attacked in the laboratory were transplanted<br />

in mid-October to a spurge stand in Saskatchewan (Table 43) to check on<br />

the ability <strong>of</strong> the larvae to survive the winter. The plants did not have time to establish<br />

themselves before winter and some were killed. In the following May one larva was<br />

found in 12 crowns examined and subsequently some oviposition girdling was found at<br />

the site but no eggs or larvae.<br />

Field collected beetles were released on three spurge stands in 1980 (Table 43) and<br />

during the summer, stems with oviposition scars were tagged and then harvested in<br />

September to determine the number <strong>of</strong> larvae that had tunnelled into the roots before<br />

winter (Table 44).<br />

The Cardston site was extremely fertile and the spurge so tall and dense that many<br />

stems with oviposition scars were certainly missed but it had by far the best ratio <strong>of</strong><br />

oviposition scars to larvae. Unfortunately the site was burnt just before winter and in the<br />

following June the spurge was sprayed with pic\oram. Thus it was hardly surprising that<br />

no sign <strong>of</strong> the beetle was found in August 1981. In contrast at Caronport winter survival


Table 44<br />

Evaluation <strong>of</strong> Control Attempts<br />

Euphorbia esula-virgata complex. 167<br />

Oviposition attempts and success by o. erythrocephala (Schr.) at three sites<br />

No. stems with No. larvae in No. larvae in<br />

Site oviposition scars roots by September stems in Septcmber<br />

Jameson. Saskatchcwan 25 1 3<br />

Caronport. Saskatchewan 93 15 2<br />

Cardston. Alberta 29 17 2<br />

was good in spite <strong>of</strong> a poor snow cover as live larvae were found in three out <strong>of</strong> four<br />

plants examined in May. About twice as many larvae were found at the end <strong>of</strong> the<br />

summer as in the previous year including two plants (untagged the year before) with<br />

larvae developing over a two year period. The colony had a radius <strong>of</strong> about 25 m in the<br />

fall <strong>of</strong> 1981. No progeny were found at the Jameson release site.<br />

In 1981, beetles were released at three widely separated sites but as in the previous<br />

year breeding success was poor (Table 43). The table indicates that the date <strong>of</strong> release had<br />

little influence on breeding success so the problem does not appear to be related to the<br />

synchrony <strong>of</strong> the plant and larval development.<br />

The larvae developed readily in the stems <strong>of</strong> spurge plants from the Jameson area<br />

grown in pots but these plants tended to have larger stems than most <strong>of</strong> those in the field.<br />

The colony at Caronport was still present in 1982 and for the first time one adult was<br />

found (a fertilized female). A search made in mid-September produced one larva which<br />

was still in the stem only a few centimeters from the oviposition point. As most <strong>of</strong> the<br />

spurge stems were dead by October it had little prospect <strong>of</strong> survival. The problem is<br />

apparently not cool summer temperatures as the Cardston site had fewer heat units than<br />

the other release sites. The stem diameter at Caronport was, however, considerably less<br />

than at Cardston (the soil is sand and too light to be cultivated) so it appears that the<br />

beetle is only likely to thrive on spurge plants growing on fertile soils.<br />

The biological control <strong>of</strong> spurge in <strong>Canada</strong> has so far been a failure. The one insect<br />

established on cypress spurge has had little impact and the possible impact <strong>of</strong> O.<br />

erythrocephala on leafy spurge will not be apparent until its population has reached its<br />

plateau. Other insects tested have failed to survive on Canadian leafy spurge although<br />

some like C. tenthrediniformis appear to have a major impact on E. esula in Europe. The<br />

initial difficulty arose from equating the E. esula <strong>of</strong> Europe with the plant called E. esula<br />

in North America. Some difficulties still remain as it is not possible to search Europe<br />

for the Canadian spp. <strong>of</strong> leafy spurge for possible agents since the precise origin and<br />

identity <strong>of</strong> this spurge is not clear. Indeed, some <strong>of</strong> it may have arisen in North America<br />

through hybridization. Thus agents have to be found by testing the acceptability <strong>of</strong> North<br />

American leafy spurge to organisms found attacking closely related European spurges<br />

(those in the section Esu/a). A number <strong>of</strong> these are already known and can be subjected to<br />

the usual screening tests to establish their specificity. A few species have already been<br />

screened and effort should be made to establish them as widely as possible.<br />

The difference in the density <strong>of</strong> spurge in European and North American stands<br />

appears to be related to the cropping pressure by a complex <strong>of</strong> insects and pathogens in<br />

Europe, although it would be nice to demonstrate this by removing the insects from a<br />

European stand. One possible explanation for the richness <strong>of</strong> the complex in Europe is


168 P. Harris<br />

Recommendations<br />

that like R. euphorbiae in Ontario each insect species is only able to utilize a small<br />

proportion <strong>of</strong> the total resource. The establishment <strong>of</strong> a complex <strong>of</strong> insects in North<br />

America large enough to reduce spurge density seems to be largely a matter <strong>of</strong> time and<br />

effort. Meantime any insect that can be established in reasonable numbers should be<br />

regarded as a success as it contributes to the cropping pressure on the weed in North<br />

America.<br />

The priorities in the biological control <strong>of</strong> leafy and cypress spurge for the immediate<br />

future are recognized as follows:<br />

(1) To complete host specificity studies on additional biological control agents in<br />

order to obtain approval for their release in North America. A reasonable target is the<br />

establishment <strong>of</strong> 10 agents in <strong>Canada</strong> since several species <strong>of</strong> spurge are probably<br />

involved.<br />

The root feeding beetles in the genus Aphthona (Chrysomelidae) are <strong>of</strong> particular<br />

interest as they are widespread and common on Euphorbia spp. in Europe. Investigations<br />

by Maw (1981) showed that A. cyparissiae Koch, A. {lava Guill., and A.<br />

czwalinae Weise would accept Canadian leafy spurge although the first two species are<br />

most common on E. cyparissias and the latter on E. esula. A. cyparissiae is likely to be<br />

the most useful for the Canadian prairies: it has the broadest ecological range <strong>of</strong> the three<br />

species, but prefers dry open sites with dense spurge and the larvae are reasonably cold<br />

tolerant. A. czwalinae will probably be restricted to moist sites such as along streams<br />

and in well wooded areas. A. f1ava prefers dry open sites but is less cold tolerant than A.<br />

cyparissias and so may be most applicable for southern Ontario or the United States.<br />

The following insects have been bred on Canadian leafy spurge in the laboratory and<br />

warrant further investigation as biological control agents. The leaf-tying moth Lobesia<br />

euphorbiana (Ferr.) (Tortricidae), the defoliating moths Minoa murinata Scop. and<br />

Bistonfiducarius Anker (Geometridae): both the former two moths in feeding tests were<br />

restricted to a few species <strong>of</strong> Euphorbia and they appear suitable for use as biological<br />

control agents. The gall midges Bayeria capitigena Bremi, Dasineura capsulae Kiett.,<br />

and D. loewi Mik. (Cecidomyiidae); the aphids Acyrthosiphum cyparissiae Koch and<br />

Aphis euphorbiae Kltb. (Aphididae) are also worthy <strong>of</strong> testing.<br />

(2) The root-feeding moth, Chamaesphecia empiformis Esp. (Sesiidae), can almost<br />

certainly be established on the tetraploid E. cyparissias in Ontario and Quebec.<br />

Renewed efforts should be made to do this.<br />

(3) Chamaesphecia tenthrediniformis (D. & S.) should be established on E. esula s.<br />

str. if this spurge is a problem anywhere in North America. It is the most effective agent<br />

on E. esula in Austria.<br />

(4) The poor survival <strong>of</strong> Oberea erythrocephala (Cerambycidae) on the Canadian<br />

prairies may be related to the low fertility <strong>of</strong> most <strong>of</strong> the release sites. Releases<br />

should be tried again at Cardston, Alberta, and on cypress spurge in Ontario. Possibly<br />

the colony at Caronport can be made to flourish and hence be a source for distribution if<br />

the site is fertilized with nitrogen.<br />

(5) The populations <strong>of</strong> the spurge hawkmoth, Ryles euphorbiae (L.) (Sphingidae),<br />

established on cypress spurge near Mulmur, Ontario, suggest that the species may have<br />

value in other places in Ontario and Quebec. The strain imported from Europe appears<br />

to have a strong ovipositional preference for cypress spurge over leafy spurge. Stock<br />

from the colony established on leafy spurge in Montana should be tried again on leafy<br />

spurge in western <strong>Canada</strong>. As summer temperature may be limiting on much <strong>of</strong> the<br />

Canadian prairies, the releases should be made first in the regions receiving the greatest<br />

number <strong>of</strong> Corn Heat Units such as around Medicine Hat, Alberta, Estevan, Saskatchewan,<br />

and southcentral Manitoba.


Blank Page<br />

170


Pest Status<br />

Background<br />

Chapter 36 171<br />

Hypericum perforatum L., St. John's-wort<br />

(Hypericaceae)<br />

P. HARRIS and M. MAW<br />

In the early 1950s, St. John's-wort was forming large stands on the grasslands in the<br />

interior <strong>of</strong> British Columbia where it became the dominant herbaceous species. A number<br />

<strong>of</strong> biological control agents were introduced (McLeod 1962, Harris & Peschken 1971) and in<br />

most areas the weed has been reduced to scattered plants or patches up to 2 m in diameter<br />

and usually 100 to 200 m apart. The use <strong>of</strong> herbicides against the weed on public and private<br />

lands practically ceased. The main agent responsible for this reduction has been the beetle<br />

Chrys<strong>of</strong>ina quadrigemina (Suffr.). In the few areas, such as the eutric brunisols on fluvial<br />

glacial gravel and morainal deposits near Cranbrook, British Columbia, where the beetle<br />

popUlation has been patchy, the weed has continued to spread. In this area, dense stands <strong>of</strong><br />

St. John's-wort occur on about 30 km! <strong>of</strong> lightly grazed Douglas fir, Pseudotsuga menziesii,<br />

forest. At present, the weed is causing little monetary loss, but if it continues to spread,<br />

extensive wildlife and cattle grazing areas in the east Kootenays are threatened.<br />

In Ontario, St. John's-wort continues to be a common weed on poor quality grasslands,<br />

on thin dry soils, on gravel, limestone, or rock; but it was removed from the noxious weed<br />

list when the Weed Control Act was amended in 1976. If the availability <strong>of</strong> an effective<br />

biological control agent contributed to its removal from the list, biological control can be<br />

considered to have been a success. The weed is localized in New Brunswick and Nova<br />

Scotia, but it is capable <strong>of</strong> forming dense stands and appears to be spreading.<br />

The background to the biological control <strong>of</strong> St. John's-wort in <strong>Canada</strong> was described by<br />

McLeod (1962) (Harris & Peschken 1971). In these accounts, the beetle C. quadrigemina<br />

was most effective in the semiarid to dry subhumid regions (Harris el af. 1969) as they had<br />

an obligatory aestival diapause which they could not enter under continuously wet conditions<br />

(Huffaker 1967). The smaller beetle, C. hyperici, was most effective in the moist<br />

subhumid to humid regions. This assessment must now be revised with the spread <strong>of</strong> C.<br />

quadrigemina onto the moist coastal plain at Agassiz, British Columbia. In the years<br />

immediately following the release <strong>of</strong> the beetles, eggs and larvae were only found on the<br />

procumbent foliage from September to May. In 1981 in a moist region <strong>of</strong> British Columbia,<br />

eggs and first instar larvae were found on 23 June on the upright stems just prior to bloom<br />

while in the drier regions the beetle was in the adult pre-aestival feeding stage.<br />

The colour forms <strong>of</strong> C. quadrigemina vary markedly between release sites in British<br />

Columbia. Pesch ken (1972) found that the prevalence <strong>of</strong> the bronze forms was correlated<br />

with the mildness <strong>of</strong> the winter. He also found several behavioural adaptations between<br />

beetles in British Columbia and those in California which were the source <strong>of</strong> the British<br />

Columbia stock.


172 P. Harris and M. Maw<br />

Releases and Recoveries<br />

Agrllus hyperici<br />

(Crotch) (Coleoptera:<br />

Buprestidae)<br />

Analtis plag/aIB (L.)<br />

(Lepidoptera:<br />

Geometridae)<br />

Aphis chloris (Koch)<br />

(Homoptera:<br />

Aphididae)<br />

Renewed attempts to establish the root boring beetle A. hyperici in the Cranbrook, British<br />

Columbia, area (Table 45) were not successful. This beetle is established at Mt. Shasta.<br />

California, and the climate <strong>of</strong> southern British Columbia is within the range <strong>of</strong> the species in<br />

Europe. The problem seems to be related to shipping stress in the newly emerged beetles,<br />

two shipments arrived dead at Cranbrook airport and no eggs or larvae were found from the<br />

release <strong>of</strong> live beetles. If further attempts are made to establish this insect, it is suggested<br />

that it should be done by transferring newly hatched larvae to plants growing at the release<br />

site.<br />

Two releases <strong>of</strong> A. plagiala were made in the Cranbrook British Columbia area and one<br />

in New Brunswick. No establishment has been reported from New Brunswick. In<br />

British Columbia, two moths were caught in September 1980, about 3 km from the<br />

release site and on 23 June 1981, when the first flowers had opened, an egg, a first instar<br />

larva, and several moths were seen. The moth was also found in 1981 near Grandforks,<br />

an area where it had been released in 1967 and not previously recovered. The moths<br />

were flying actively between the widely scattered clumps <strong>of</strong> the weed and had survived<br />

in spite <strong>of</strong> strong populations <strong>of</strong> C. quadrigemina. Many larvae were found at Grandforks in<br />

late July and a few near Cranbrook in late July and early October (K. Williams 1981<br />

personal communication). The colony at Agassiz, British Columbia, in 1981 caused<br />

browning and defoliation <strong>of</strong> H. perforalum in patches (L. Sigurgeirson 1981 personal<br />

communication).<br />

The A. plagiala larvae released in 1977 at North Tay Creek, New Brunswick, may have<br />

been killed by insecticide drift from aerial spray <strong>of</strong> the adjacent forest for spruce<br />

budworm (D. Finnamore 1981 personal communication).<br />

(a)Ecology<br />

The sexual generation <strong>of</strong> the aphid A. chloris appears in the fall (October in Cranbrook)<br />

in response to lower temperatures and shortened day length. During this period and in<br />

early November. an average <strong>of</strong> four (maximum <strong>of</strong> eight spread over a month) yellow<br />

eggs, which tum shiny black, are laid per female at the base <strong>of</strong> the flowering stems or on<br />

the procumbent foliage. Hatching occurs in April or May giving rise to a generation <strong>of</strong><br />

apterous females. These mature in about a month and several generations <strong>of</strong> apterous<br />

viviparous females are produced. During midsummer, when colonies become crowded.<br />

alate viviparous females develop and disperse to found new colonies.<br />

A. chloris has a wide geographic range. It is found in Sweden near the northern limit <strong>of</strong><br />

St. John's-wort (Johansson 1962). in Poland (Czechowskii 1975), in Israel (Neser. 1973,<br />

personal communication), in Britian in cool and damp conditions, and in the arid south<br />

<strong>of</strong> France, from sea level to 1500 m in the mountains (Wilson 1943).<br />

In the south <strong>of</strong> France and at Cranbrook, the aphids in summer are found near the tip<br />

<strong>of</strong> the shoots or beneath the flowers. In contrast, in England, northern France. and in the<br />

Alps, they are normally found at the base <strong>of</strong> the stems. Wilson (1943) suggested that the<br />

feeding site depends on conditions <strong>of</strong> heat and moisture. If so, the ability <strong>of</strong> the aphid to


Chrysolina hypericJ<br />

(Forester)<br />

(Coleoptera:<br />

Chrysomelidae)<br />

Ilypericum perfOrUlIlm L.. 173<br />

adjust in this manner bodes well for its survival in the varied conditions at Cranbrook.<br />

British Columbia.<br />

In France, A. chloris can be heavily parasitized by Ap/,idius cardui Marsh. (Braconidae)<br />

and Aphidencyrtus aphidiverus Mayr. (Encyrtidae), but its reproductive capacity is<br />

such that it remains a common species (Wilson 1943). The stocks <strong>of</strong> A. chloris received<br />

from the Rhine Valley and Alsace contained Lysiphlebus faborum (Marsh.) (Braconidae).<br />

This is an aggressive parasitoid and if left unchecked can destroy a laboratory colony.<br />

Although A. chloris will attack various species <strong>of</strong> Hypericum, the damage done to the<br />

plant or its acceptance by the aphid varies with the species. For example, at the end <strong>of</strong> 5<br />

days the number <strong>of</strong> <strong>of</strong>fspring from two mature apterous females on H. perforatum. H.<br />

rhodopaeum, H. calyeinum, and H. densiflorum were 37, 10, 7, and 0, respectively.<br />

Large colonies <strong>of</strong> A. chloris consistently killed vigorous plants <strong>of</strong> H. perforatum in<br />

about a month, while H. rhodopaeum lived for over seven months and put on new<br />

growth in spite <strong>of</strong> a heavy aphid population.<br />

(b) Releases<br />

Two releases <strong>of</strong> A. chloris were made in the Cranbrook, British Columbia, area (Table<br />

45). Infested H. perforalum plants were transplanted into the field. In 1979 this was done<br />

during a period <strong>of</strong> drought: the transplanted plants died and most <strong>of</strong> the surrounding H.<br />

perforatum had lost their foliage. No evidence <strong>of</strong> aphid survival was found. In 1980<br />

infested plants were transplanted earlier in the spring and the aphids moved readily onto<br />

the surrounding plants. By early November all <strong>of</strong> the rosettes sampled in a 20 m radius at<br />

the two release sites had a few fundatrigeniae and/or eggs. By late June 1981 following a<br />

wet and cool spring, scattered plants within a radius <strong>of</strong> 50 m <strong>of</strong> the release point had<br />

small colonies <strong>of</strong> A. chloris near the flower buds. By far the strongest colony <strong>of</strong> the aphid<br />

was on the site least favoured by C. quadrigemina.<br />

British Columbia was used as a source for obtaining stock for release in Ontario, Nova<br />

Scotia, and New Brunswick (Table 45).<br />

In Ontario, C. hyperici became established at the release site near Picton. Overwintering<br />

survival was 63% for adults and 31% for eggs compared to 19% and 6%<br />

respectively, for C. quadrigemina (Harris & Peschken 1974). Nevertheless C. quadrigemina<br />

rapidly became the most common species, although both species were still<br />

present in 1979 (13 ddand4 22 C. quadrigemina:4 ddand 322 C. hypericiin the sample<br />

collected).<br />

The releases in Nova Scotia contained a few individuals <strong>of</strong> C. quadrigemina collected<br />

with the C. hyperici from Fruitvale, British Columbia. Both species became established<br />

but by 1979 C. hyperici had become numerous enough at one site in the Annapolis Valley<br />

to generate one query about the identity <strong>of</strong> the beetles sitting on the side <strong>of</strong> a house (H.<br />

Specht, 1979, personal communication). This was about 30 km from the release site.<br />

According to M. Neary (1979, personal communication), the original release site has<br />

been cleared <strong>of</strong> the weed and it has been substantially reduced in the Kingston-Auburn<br />

area. C. hyperici is also established in the Dundurn and Fenwick regions. At Dundurn<br />

between 1970-72 the number <strong>of</strong> flowering stems declined from 5.8 to 3.4/m2. This is a<br />

more gradual decline than has been usual with C. quadrigemina in British Columbia.<br />

C. hyperici released at Tay Creek, New Brunswick, were well established in 1977, but<br />

were not numerous enough to control the weed. A problem with the release made in 1975<br />

is that all the individuals examined were female. Apparently the females continued their


174 P. Harris and M. Maw<br />

Table 45 Open releases and recoveries <strong>of</strong> insects against Hypericum per/oralum L.<br />

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

Release site Year Origin Number Recovery<br />

Chrysolina hyperici<br />

Pictou Co., Nova Scotia 1969 British Columbia 2 100 adults 1970<br />

Prince Edward Co., Ontario 1969 British Columbia 70 adults 1970<br />

Annapolis Co., Nova Scotia 1969 British Columbia 1320 adults 1979<br />

Prince Edward Co., Ontario 1970 British Columbia 182 adults 1970<br />

Tay Creek, New Brunswick 1971 British Columbia 1 150 adults 1972<br />

Lime Kiln, New Brunswick 1971 British Columbia 1 150 adults<br />

Tay Creek, New Brunswick 1975 British Columbia 496 adults<br />

Chrysolina quadrigemina<br />

Prince Edward Co., Ontario 1969 British Columbia 182 adults 1970<br />

Sidney Twp., Ontario 1971 British Columbia 15 adults<br />

1971 California 139 adults<br />

Tay Creek, New Brunswick 1975 British Columbia 128 adUlts} Not found<br />

Tay Creek, New Brunswick 1977 British Columbia 55 adults in 1981<br />

Guelph, Ontario 1979 Ontario 40 adults· 7000 collected<br />

in 1981<br />

4 sites near Guelph, Ontario 1979 Ontario 560 adults No recovery<br />

34 sites in and around<br />

Guelph, Ontario 1981 Ontario 7000 adults<br />

Agrilus hyperici<br />

Elko, British Columbia 1977 California 56 adults No recovery<br />

up to 1981<br />

Aphis chloris<br />

Elko, British Columbia 1979 Germany 300 plus Did not<br />

survive<br />

Elko, British Columbia 1980 Germany 150000 1981<br />

Anailis plagiata<br />

Grandforks, British Columbia1967 Europe 875 larvae 1981<br />

Westbridge, British Columbia 1967 Europe 500 larvae<br />

Tay Creek, New Brunswick 1977 Switzerland 715 adults Not found<br />

in 1978<br />

Elko, British Columbia 1977 Switzerland 3648larvae<br />

Agassiz, British Columbia 1977 Switzerland 476 larvae 1981<br />

Elko, British Columbia 1980 France 169 larvae 1981**<br />

• Sample examined contained 17 C. quadrigemina: 7 C. hyperici<br />

.. In view <strong>of</strong> the small population in 1981, it is assumed that establishment occurred<br />

from the 1980 releases.<br />

pre-aestival feeding on the flowering plants for longer than the males, so that mass<br />

collecting produced a badly distorted sex ratio. It is not known whether other releases<br />

were adversely affected in this manner. The release site at North Tay Creek, New<br />

Brunswick, has been subject to insecticide drift from aerial treatments <strong>of</strong> the forest<br />

against spruce budworm. In 1978, 1979, and 1980, a buffer strip <strong>of</strong> two miles was left<br />

between the treated forest and the release field and beetle density increased noticeably


176 P. Hnrris nnd M. MlIw<br />

Recommendations<br />

beetle, C. hyperici, was needed to control the weed in moist sites. C. quadrigemina has now<br />

become numerous at the moist sites <strong>of</strong> Fruitvale and Agassiz, British Columbia, although<br />

C. hyperici has persisted and spread to other sites in the province (K. Williams 1981<br />

personal communication). However, in spite <strong>of</strong> the release <strong>of</strong> pioneer moist-adapted<br />

individuals <strong>of</strong> C. qlladrigemina with C. hyperici in the maritimes, the latter species has<br />

become dominant and C. quadrigemina was not present in the samples examined. C. hyperici<br />

is only slightly smaller than C. quadrigemina and the two species feed in a similar manner on<br />

H. perforattun, so it is surprising that they coexist in the same habitats. It is not known what<br />

occurred to allow C. quadrigemina to colonize moist habitats and invade the thriving colony<br />

<strong>of</strong> C. hyperici at Fruitvale. The adaption has not extended to the maritimes and again the<br />

reasons are unknown. From a purely pragmatic purpose <strong>of</strong> controlling H. perforatum the<br />

important thing is that one species or another will reduce the weed to a low density in most<br />

places in <strong>Canada</strong>. For example, at the release site in Ontario, the weed declined to only 0.2%<br />

<strong>of</strong> its former density.<br />

A strong colony <strong>of</strong> C. quadrigemina with some C. hyperici was established near<br />

Guelph, Ontario, with stock from the Picton, Ontario, colony. A total <strong>of</strong> 7000 beetles<br />

collected from the Guelph colony before its destruction in 1981 was distributed to 34<br />

other sites in and around the city. By 1983 there should be massive populations <strong>of</strong> the<br />

beetle in the Guelph area. These could be collected at little cost and released at other H.<br />

perforatum sites throughout Ontario. Some <strong>of</strong> the releases would fail (two out <strong>of</strong> five<br />

release sites <strong>of</strong> 1979 at Guelph were destroyed) but the beetle would distribute itself<br />

from the survivors. The pilot studies at Picton, Ontario as well as those in British<br />

Columbia indicate that the prevalence <strong>of</strong> H. perforatum would be greatly decreased by<br />

such a programme with a corresponding increase in forage for livestock or wildlife.<br />

The apparently few sites where Chrysolina spp. have been ineffective are on nutrientpoor<br />

soils and in insecticide-drift areas. No insects are likely to be effective for biological<br />

control if they receive an insecticide spray while they are exposed on their host<br />

plant. It might be possible to find a root-boring insect that was inside the plant during the<br />

spruce budworm spraying season; however, in view <strong>of</strong> the public outcry, the spray<br />

programme is likely to become increasingly circumspect, so the problem for C. hyperici<br />

should diminish.<br />

More investigations are needed to determine whether the effectiveness <strong>of</strong> the Chryso­<br />

Iina spp. can be increased on nutrient-poor soils; however, as the moth A. plagiata is<br />

increasing on some <strong>of</strong> the areas not favored by the beetle, we already may have part <strong>of</strong><br />

the solution. Aphis chloris was chosen as a biological control agent as the presence <strong>of</strong><br />

symbionts in aphids <strong>of</strong>ten enables them to survive on plants that lack one or more<br />

essential amino acids or B vitamins. However, its effectiveness as a biological control<br />

agent for H. perforattun remains to be seen.<br />

(1) C. quadrigemina and/or C. hyperici have shown that they are able to reduce H.<br />

perforatum to well below the economic threshold in most regions <strong>of</strong> <strong>Canada</strong>. Any<br />

province with H. perforaltun on their noxious weed list should distribute the beetles or,<br />

if the importance <strong>of</strong> the weed does not justify this small effort, remove the weed from the<br />

list.<br />

(2) Provinces where the weed is being satisfactorily controlled should remove H.<br />

perforalum from their noxious weed list as mowing or spraying operations against the<br />

weed are a waste and not likely to assist its biological control.<br />

(3) Further study is required to determine the role <strong>of</strong> Anaitis plagiata and/or Aphis<br />

chloris.<br />

(4) If further attempts are made to establish Agrilus hyperici, these should be done<br />

by transferring the young larvae to plants in the field.


Blank Page<br />

178


Pest Status<br />

90<br />

80<br />

70<br />

I/)<br />

-u<br />

ctI<br />

"- 50<br />

-I/)<br />

.0<br />

ctI<br />

"t:I<br />

Q) 50<br />

-ctI<br />

::::l<br />

E<br />

::J<br />

u 40<br />

u<br />

«<br />

30<br />

20<br />

10<br />

Chapter 37<br />

Linaria vulgaris Miller, Yellow Toadflax<br />

and L. dalmatica (L.) Mill., Broad-leaved<br />

Toadflax (Seroph ulariaeeae )<br />

P. HARRIS<br />

The importance <strong>of</strong> Linaria vulgaris Mill. in western <strong>Canada</strong> has declined since the early<br />

1950s so that now, as in eastern <strong>Canada</strong>, it is a conspicuous but relatively unimportant<br />

weed. In a study in the Peace River area <strong>of</strong> Alberta, Darwent et al. (1975) found that in<br />

• •<br />

54 56 58 60 62 64 66 68 70 72 74 76 78<br />

Year<br />

Fig. R. Cumulative number <strong>of</strong> abstracts on the control <strong>of</strong> toadnax with hcrbidcs in the Research Report <strong>of</strong> the Expert<br />

CommiUec on Weeds (<strong>Canada</strong> Western Section)<br />

179


180 P. Harris<br />

Background<br />

Table 47<br />

most areas it was less abundant in the early 1970s than it had been in 1956 and that<br />

toadflax was not a serious problem in cereal crops. In creeping red fescue, 180 toadflax<br />

stems/m 2 reduced the yield by a third; but in most areas there were 20 or less stems/m 2<br />

with no measurable effect on yield. Most <strong>of</strong> the dense stands <strong>of</strong> the weed were on sites<br />

that had been cleared and abandoned or were cropped infrequently.<br />

In Saskatchewan, the weed is still dense on uncultivated sandy soils from North<br />

Battleford west to Alberta. In other regions <strong>of</strong> the province it has disappeared or greatly<br />

declined since the 1950s.<br />

The decrease in the importance <strong>of</strong> the weed is reflected by the decline in the number <strong>of</strong><br />

abstracts relating to its chemical control in the Research Report <strong>of</strong> the Expert Committee on<br />

Weeds (<strong>Canada</strong> Western Section) (Fig. 8). The decrease <strong>of</strong> the weed has not however<br />

changed its status under noxious weed legislation. For example in Saskatchewan provincial<br />

subsidies are still paid to municipalities for its control.<br />

L. dalmatica (L.) is considerably more persistent that L. vulgaris, and it appears to<br />

exclude most other herbaceous species from the stand. Stands are local and the total<br />

area occupied does not appear to be large. Thus in the absence <strong>of</strong> documented evidence<br />

it does not appear to <strong>of</strong>fer a serious threat.<br />

A more detailed history <strong>of</strong> the biological control <strong>of</strong> toadflax is given by Harris &<br />

Carder (1971). The decline <strong>of</strong> the weed in western <strong>Canada</strong> coincided with the appearance<br />

<strong>of</strong> the flower-feeding beetle Brachypterolus pulicarius (L.). This European<br />

beetle was present in eastern <strong>Canada</strong> and spread to the west. In addition the seedfeeding<br />

weevil Gymnaetron antirrhinii (Payk.), another European insect present in the<br />

east, was released in western <strong>Canada</strong> in 1957. Both insects are now present in regions <strong>of</strong><br />

western <strong>Canada</strong> where toadflax is common.<br />

In 1962, a European defoliating moth, Calophasia lunula (Hfn.), was first released in<br />

<strong>Canada</strong>, and it became established on the railway right-<strong>of</strong>-way at Belleville, Ontario in<br />

1966. The moth has two overlapping generations and lays single well distributed eggs.<br />

The first two larval instars are inconspicuous and frequently feed inside the flowers, but<br />

the third to fifth instars feed externally on the foliage and are easily seen. Samples taken<br />

along the railway right-<strong>of</strong>-way (Table 47) indicate that larvae were present from late May<br />

to early September, usually at a density <strong>of</strong> 2-6 larvae/100 stems regardless <strong>of</strong> the<br />

toadflax density. The density <strong>of</strong> the toadflax increased greatly following soil disturbance<br />

in the previous year. The overall level <strong>of</strong> damage was not impressive with less than 20"10<br />

<strong>of</strong> the stems being defoliated.<br />

Density <strong>of</strong> C. lunula and toadflax, L. vulgaris Mill. and L. dalmatica (L.) Mill .• along<br />

the railway at Belleville. Ontario.<br />

No. 3rd-5th instar<br />

Date No. samples· No. Toadflax stemsfm Z C. lunula larvae/m 2<br />

2 Jun 1968 26 85.7 2.0<br />

8 Aug 1968 19 89.1 3.8<br />

3 Sep 1969 158 6.6 0.05<br />

18 Jun 1970 52 7.2 0.31<br />

13 Aug 1970 205 0.98<br />

15 Aug 1971 31 7.7 0.13<br />

• (0.25 m 2 ) samples taken at 5 pace intervals


Table 48<br />

Releases and Recoveries<br />

Evaluation <strong>of</strong> Control Attempts<br />

Linaria \·ulgaris Miller. 181<br />

The area has not been surveyed since 1972, but the moth has continued to spread; it<br />

was recovered from light traps at Ottawa, Ontario in 1977, a distance <strong>of</strong> over 300 km<br />

from the release site (E. Monroe, 1978, personal communication). Thus C. lunula is now<br />

probably present in most <strong>of</strong> southern Ontario. According to I. Jones (1978, personal<br />

communication), larvae were scarce in the Ottawa area in 1980: less than half a dozen<br />

were seen in several organized and many casual searches. In 1981, there was a large<br />

increase in the population, larvae were seen from early June to mid-October and along<br />

the railway track 47 to 57 larvae could be collected in an hour. Defoliated stems were<br />

evident, and some small stands were completely defoliated. Both the larvae and defoliation<br />

were associated with stunted toadflax plants and they rarely occurred in adjacent<br />

lush stands.<br />

Op'en releases and recoveries <strong>of</strong> Calophasia [wUlla against L. vulgari.v Mill. and L.<br />

dalmatica (L.) Mill.<br />

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

Weed Species Province Year Origin Number Recovery<br />

L. vulgaris New Brunswick 1971 Ontario 1565 larvae<br />

& 136 pupae<br />

L. vulgaris Saskatchewan 1973 Ontario 2725 larvae<br />

L. dalmatica British Columbia 1969 Ontario 200 pupae<br />

L. dalmatica British Columbia 1971 Ontario 1180 larvae<br />

Four attempts were made to establish Calophasia lunula in regions <strong>of</strong> <strong>Canada</strong> outside<br />

Ontario and all were unsuccessful (Table 48). The release site in Saskatchewan was<br />

flooded in the year following the release. The larvae released on Dalmation toadflax in<br />

British Columbia disappeared within days, apparently from ant predation. Survival<br />

was not followed closely in New Brunswick but no evidence <strong>of</strong> establishment was<br />

found in the following year.<br />

The moth, C. lunula, is obviously well established in Ontario and it should be possible<br />

to establish it in the rest <strong>of</strong> <strong>Canada</strong>. C. lunula has not defoliated enough L. vulgaris to<br />

control the weed on its own, but it has added to the stress inflicted by the now<br />

ubiquitous beetles, G. antirrhinii and B. pulicarius. The possibility that it may<br />

sometimes inflict more damage is indicated by the populations in Ottawa in 1981 which<br />

seem to be about ten times the density at Belleville prior to 1972.<br />

C. lunula favors the stunted toadflax stands that are found on dry soils. It is these<br />

stands that are persistent as edaphic conditions that permit the formation <strong>of</strong> tall lush<br />

stands also lead to the displacement <strong>of</strong> toadflax by grass and other vegetation. Thus C.<br />

lunula concentrates its attack where the weed is most persistent.<br />

The combination <strong>of</strong> the three toadflax insects can control L. vulgaris at below the<br />

economic threshold in most parts <strong>of</strong> <strong>Canada</strong>. Indeed in most regions this has been<br />

achieved by the two beetles alone. An exception is North Battleford, Saskatchewan,


182 P. Harris<br />

Recommendations<br />

Acknowledgements<br />

Literature Cited<br />

where there are extensive dense stands <strong>of</strong> toad flax. Based on the results from Ontario,<br />

it is doubtful whether the establishment <strong>of</strong> C. lunula in the area would solve the<br />

problem. Whether it is worth screening and establishing a fourth agent for the control<br />

<strong>of</strong> toadflax around North Battleford needs to be determined by a cost-benefit study as<br />

opinions vary about the importance <strong>of</strong> toadflax to the region.<br />

Surveys for the biological control agents <strong>of</strong> toadflax and their host specificity testing<br />

should not be done by <strong>Canada</strong> unless it can be shown that the potential benefits are<br />

large enough to justify the costs <strong>of</strong> biological control. However, the costs to <strong>Canada</strong><br />

for biological control would be minimal if the tests were done by other countries. In<br />

this event the control agent should be imported as there would be benefits to individual<br />

land holders from the biological control <strong>of</strong> Dalmation and common toadflax. The moth<br />

C. lunula can almost certainly be established outside Ontario.<br />

I am grateful to Mr Ian Jones for the observations on C. lunula in the Ottawa area in<br />

1980 and 1981.<br />

Darwent, A.L.; Labay, W.; Yarish, W.; Harris, P. (1975) Distribution and importance in northwestern Alberta <strong>of</strong> loadflax and its insect<br />

enemies. Catuldian Journal 01 Piant Science 55, 157-162.<br />

Harris, P.; Carder. A.C. (1971) Linaria vulgaris Mill .• yellow loadfiax and L. dalmaJica (L.) Mill .• broad-leaved toadflax (Scrophulariaccae). In:<br />

Biological control programmes against insects and weeds in <strong>Canada</strong> 1959-1968. Commonwealth<br />

inslilute 01 Biological Control Technical Communication 4. 94-97.


Chapter 38<br />

Opuntia polyacantha Haworth, Plains<br />

Prickly-pear Cactus (Cactaceae)<br />

M.G. MAW<br />

There are three species <strong>of</strong> cacti on the Canadian Prairies: the prickly-pears, Opuntia<br />

polyacantha Haw. and O. tragi/is (Nutt.) Haw., and the pincushion or ball cactus,<br />

Mammillaria vivipara (Nutt.) Haw. Mammillaria is not considered to be a problem.<br />

In the western provinces cacti are found on dry hillsides and flatlands throughout the<br />

more arid parts <strong>of</strong> the brown soils zone. These areas have some grazing capacity, but<br />

have limitations that make improvement and cultivation impractical. For example,<br />

when the surface is disturbed the soils are likely to erode and blow easily.<br />

Periodically, especially during droughts, prickly-pear becomes a concern. Ranchers<br />

then worry that it will displace valuable grass, use scarce moisture, and cause injury to<br />

stock. It is generally thought that overgrazing increases cactus growth and spread.<br />

Certainly, severely overgrazed grasslands greatly facilitate the spread <strong>of</strong> cacti for when<br />

soils have been denuded, cactus pads broken from clumps come into contact with bare<br />

soil, root, and establish new clumps. On the other hand, Bement's (1968) study <strong>of</strong> heavy,<br />

moderate, and light pasture use showed that while populations <strong>of</strong> cactus doubled in area<br />

in 28 years, it never exceeded 2.4% <strong>of</strong> land area. Removal <strong>of</strong> cactus did not increase<br />

forage production significantly but did make it more available for grazing. In lightly<br />

grazed pastures, grasses tended to mask cactus while under heavy grazing, cactus was<br />

more conspicuous. Farmers in the Avonlea area <strong>of</strong> Saskatchewan maintain that cactus<br />

has not changed much in the past 20 years and where it has increased, it was through<br />

careless efforts to break up the clumps with harrows (personal interviews).<br />

The value <strong>of</strong> cactus in the North American ecosystem is not always appreciated.<br />

During drought and under heavy grazing, cactus reduces soil erosion. Clumps <strong>of</strong> cactus<br />

conserve moisture by holding snow, and after summer showers the soil within the<br />

clumps remains moist for several hours longer than soil outside them. Clumps also<br />

protect desirable plants from grazing so that seeds can be produced to repopulate the<br />

range after drought or heavy grazing (Houston 1963).<br />

Prickly-pear is a food for at least 44 species <strong>of</strong> birds and mammals. For example, seeds<br />

may account for 65% <strong>of</strong> the diet <strong>of</strong> the Harris ground squirrel and up to 5.0% <strong>of</strong> the<br />

browse <strong>of</strong> deer and antelope (Martin et al. 1961).<br />

The earliest record <strong>of</strong> an insect attacking cactus dates back several hundred years,<br />

and there is indication that cultivation <strong>of</strong> the cochineal, Dactylopius coccus Costa, was<br />

conducted in Mexico and Central America for many centuries before the voyage <strong>of</strong><br />

Columbus (Mann 1969).<br />

The first comprehensive study <strong>of</strong> insects <strong>of</strong> Cactaceae in the United States was done<br />

by Hunter et al. (1912). This work described the life history and habits <strong>of</strong> various<br />

species, summarized previous information. and provided an extensive and useful bibliography.<br />

Surveys for insects associated with O. polyacantha and O. /ragilis in Saskatchewan<br />

and Alberta were made from 1974 to 1979 (Maw & Molloy 1980). Although only 20<br />

species were found. three, Chelinidea vittiger Uhler, Dacty/opius con/usus (Ckll.), and<br />

Melitara dentata (Grote), were numerous enough to exert considerable pressure on the<br />

cactus. D. con/usus and M. dentala can be very destructive and each year destroy about<br />

10% <strong>of</strong> the pads (Maw & Molloy 1980)<br />

Two species <strong>of</strong> Lepidoptera. Cactob/aslis doddi Heinrich and C. bucyrus Dyar. were<br />

183


184 M. G. Maw<br />

Discussion<br />

Recommendations<br />

Literature Cited<br />

obtained from the high altitude valleys <strong>of</strong> Jujuy province in Northern Argentina for<br />

screening at Regina. The two species fed on local prickly-pear, but it was necessary to<br />

injure the pads before the larvae would enter to feed. Once inside, the larvae completely<br />

hollowed the pads, but then would not leave them to attack new ones. Either the local<br />

cactus was unsuitable or the insects were diseased for both species became weak and the<br />

colonies died out after a few generations in the laboratory.<br />

No introductions <strong>of</strong> cactus insects are considered, for any harm or inconvenience<br />

caused by our prairie cactus is balanced by benefits from it. The native biotic agents<br />

exert considerable pressure on the cactus and our native grasses, when not overgrazed<br />

are well able to compete.<br />

Any release <strong>of</strong> biological control agents to control prickly-pear would be vigorously<br />

opposed by both the United States and Mexico, because Opuntia spp. are used as<br />

emergency fodder for cattle and as human food. In addition, at least one species <strong>of</strong><br />

Opuntia is listed as endangered and 12 as threatened and still others are valued as<br />

ornamentals throughout the American southwest.<br />

(I) Introduction <strong>of</strong> biological control agents for Opuntia spp. in <strong>Canada</strong> should not<br />

be pursued.<br />

(2) Any control <strong>of</strong> cactus in <strong>Canada</strong> should be approached as a range management<br />

problem.<br />

Bement, R.E. (1968) Plains prickly-pear; relation to grazing intensity and blue grama yield on Central Great Plains. Journal <strong>of</strong> Range<br />

Matulgement 19, 83-86.<br />

Houston, W.R. (1963) Plains prickly-pear, weather, and grazing in the Northern Great Plains. Ecology 44, 569-574.<br />

Hunter, W.O.; Pratt, F.C.; Mitchell, J.D. (1912) The principle cactus insects <strong>of</strong> the United States. USDA Bureau <strong>of</strong> Entomology 113. 1-71.<br />

Mann, J. (1969) Cactus-feeding insects and mites. Smithsonian Institute Bulletin 256, 158 pp.<br />

Martin, A.C.; Zim, H.S.; Nelson, A.L. (1961) American wildlife and plants. A Guide to wildlife food habits. New York, Dover Publications<br />

Inc., 500 pp.<br />

Maw, M.O.; Molloy, M.M. (1980) Prickly-pear cactus on the Canadian prairies. Blue Jay 38, 208-211.


Pest Status<br />

Chapter 39<br />

Rhamnus cathartica L., Common or<br />

European Buckthorn (Rhamnaceae)<br />

M.G. MAW<br />

Common or European buckthorn, Rhamnw calhartica L., is the alternate host <strong>of</strong> crown<br />

rust or leaf rust <strong>of</strong> oats, Puccinia coronala Cda. f. sp. avenae Eriks.<br />

Considerable amounts <strong>of</strong> urediospore inoculum <strong>of</strong> crown rust is blown northwards<br />

every year from the United States and has nothing to do with the presence <strong>of</strong> buckthorn<br />

in <strong>Canada</strong>. Moreover, in some years, crown rust is so widespread and severe, such as in<br />

1980, that the presence or absence <strong>of</strong> buckthorn is probably not important (R. V. Clark,<br />

1980, personal communication) (Hanson & Grau 1979, Peturson 1954). Nevertheless,<br />

buckthorn is responsible for virtually all <strong>of</strong> the early infection by crown rust in the<br />

northern United States and in <strong>Canada</strong> (Hanson & Grau 1979), infecting oats from two<br />

weeks (Harder & McKenzie 1974) to a month (Harder 1975) before the arrival <strong>of</strong> the<br />

urediospores from the south.<br />

Buckthorn leaves are the site <strong>of</strong> spermagonial and the aecial stages <strong>of</strong> crown rust <strong>of</strong><br />

oats. This is a potential source <strong>of</strong> new pathogenic variation because hybridization and<br />

recombination occur in the pathogen on buckthorn (Simons el al. 1979). The greater<br />

diversity <strong>of</strong> rare races in the east, as opposed to the west, may be attributed to the<br />

presence <strong>of</strong> buckthorn (Fleischmann el al. 1963). Although there are a number <strong>of</strong> species<br />

<strong>of</strong> Rhamnw, both native and introduced, in North America, <strong>of</strong> those susceptible to P.<br />

coronala f. sp. avenae, only R. cathartica occurs in any significant numbers in<br />

<strong>Canada</strong> (Peturson 1954).<br />

It is difficult to assess overall losses due to crown rust <strong>of</strong> oats. Unless there is a general<br />

epidemic <strong>of</strong> the rust, the infection is generally extremely variable. In Ontario, in 1977,<br />

infections ranged from a trace in some fields to nearly complete destruction <strong>of</strong> the oat<br />

crop in others (Harder 1978). Yield reductions are negligible if crown rust development<br />

does not precede heading, especially in relatively dry seasons. A late attack can,<br />

however, cause losses <strong>of</strong> 763-1029 kg per ha if the season is delayed by cool and wet<br />

weather (Fleischmann 1968). When there are few late fields <strong>of</strong> oats, severe infections <strong>of</strong><br />

30-90% may cause little damage (Fleischmann 1963).<br />

In Manitoba, estimates <strong>of</strong> combined crown and stem rust damage in 1970 indicated a<br />

loss in excess <strong>of</strong> 154 000 tonnes but less than half <strong>of</strong> it was attributed to crown rust<br />

(Martens 1971). In 1969, in Manitoba, losses to crown rust were estimated at 108 ()()()<br />

tonnes and losses in individual fields ranged from 5-30% (Fleischmann 1969).<br />

It is only when there is a minimal urediospore shower that the losses due to crown rust<br />

associated with buckthorn can be determined. Buckthorn is responsible for local and<br />

persistent infestations <strong>of</strong> the rust wherever buckthorn and oats are in close proximity<br />

and a single large buckthorn is sufficient to generate a moderately severe infection in an<br />

adjacent field (Harder 1975, 1978). The distribution <strong>of</strong> crown rust in Ontario in 1977<br />

indicates that buckthorn was responsible in that year for most <strong>of</strong> the rust infection <strong>of</strong><br />

oats. The very light infection occurring in the absence <strong>of</strong> buckthorn suggests there was<br />

little influx <strong>of</strong> spores from elsewhere (Harder 1978).<br />

Crown rust adversely affects yield, quality, weight, and protein content <strong>of</strong> oat seed.<br />

Light infections cause minor reductions, but infections <strong>of</strong> 68-80% cause yield<br />

reductions <strong>of</strong> 19.2 - 27 .8%, kernal weight <strong>of</strong> 12.3 -16.1 % , and a grade reduction <strong>of</strong> about<br />

one grade (Peturson 1952).<br />

185


186 M.G. Maw<br />

Background R. cathartica was introduced into North America as a hedge and shelterbelt shrub or as<br />

an ornamental in parks and gardens. It varies considerably in size, ranging upwards <strong>of</strong> 10<br />

m high under favourable conditions but is more usually less than half that height. It is<br />

very bushy, <strong>of</strong>ten many stemmed, and the shoots and branches terminate in a sharp<br />

spine. The leaves are alternate, smooth, and elliptical with toothed margins. The flowers<br />

are in the axils <strong>of</strong> the leaves, small, greenish, and dioecious. The mature fruit is a bluish<br />

to nearly black drupe, bitter to the taste, and about 8 mm in diameter. The seed is 3-4<br />

grooved.<br />

Reproduction is entirely by seeds which germinate 90-100% when the fruit is just<br />

ripe. The plant regenerates quickly from cut stems and after burning and grazing<br />

(Montgomery 1956, Mulligan 1952, Rydberg 1932).<br />

(a) Habitat<br />

In eastern <strong>Canada</strong>, buckthorn is found along fence rows, along roadsides, in open<br />

woods, and along the gently sloping banks <strong>of</strong> lakes, rivers, and streams, it is <strong>of</strong>ten seen in<br />

solid masses along every fence row in many locations, but apparently is not found to any<br />

extent in pastures. Plants <strong>of</strong> about 2.5 m can be found in wood openings with many<br />

smaller plants and seedlings in the more shaded locations. Where the trees are removed,<br />

these smaller plants grow rapidly forming a forest <strong>of</strong> buckthorn. It tends to be more<br />

commonly on soils <strong>of</strong> high moisture content than on light, dry soils (Mulligan 1952).<br />

In Manitoba, there are relatively few areas outside <strong>of</strong> the larger towns and cities where<br />

buckthorn is found (Peturson 1954). It is concentrated in Winnipeg, Brandon, and near<br />

Macdonald, generally as park and hedgerow plantings (Bassett 1958), and in a sheltered<br />

ravine near Morden (Harder 1975).<br />

In England, it is found chiefly on alkaline peat and limestone soils (Godwin 1943).<br />

(b) Geographic range<br />

In England, it is found in the midland, southern and eastern counties, but is absent<br />

from Cornwall, North Devon, Northumberland, and western counties <strong>of</strong> Wales.<br />

In Ireland, it is present in a wide belt across the central Irish plain, but absent from the<br />

southern counties and absent or rare in the northern.<br />

It is doubtfully native in Scotland, except possibly in Dumfries.<br />

Its range extends through the greater part <strong>of</strong> Europe to 60" 48' in Norway and 61 0 41' in<br />

Sweden, in southern Finland, Esthonia, and across Russia into western Asia, Afghanistan,<br />

and Turkestan.<br />

It is present in southern Europe to middle and eastern Spain and Macedonia, and<br />

sparingly at high altitudes in Morocco and Algeria, but absent in the Balearics, Corsica,<br />

and Sardinia, also from western Siberia, Transcaucasia to Altai (Godwin 1943).<br />

In North America, R. cathartica is found in the New England states, Virginia,<br />

westward to Illinois, Missouri, Wisconsin, Minnesota (Fernald 1950), Quebec, Ontario<br />

Manitoba, and to a limited extent in Saskatchewan, Alberta, and the Maritimes.<br />

(c) Uses<br />

R. cathartica is a useful hedge and shelterbelt shrub and as such has been extensively<br />

planted in Ontario and Quc!bec and to a lesser extent in other provinces. It is also used as<br />

an ornamental in gardens and parks.


Discussion<br />

Rhamnus cathartica L.. 187<br />

The fruit is eaten by birds but usually only as a last resort. For example. robins will<br />

take them sparingly when no other food is available in the early spring. It is claimed that<br />

emodin in the green fruits prevents premature predation by birds (Trial & Dimond 1979).<br />

Perhaps late retention <strong>of</strong> fruit well into the spring attests to this.<br />

Several species <strong>of</strong> birds are cited as feeding on the fruit in Germany and England. and<br />

mice have been seen to eat the seed from dried fallen fruit.<br />

(d) Control<br />

Buckthorn is so prevalent in Ontario. especially in the eastern counties. and the bush is<br />

<strong>of</strong> such a large size that it is an almost impossible task to eradicate it by spraying. cutting,<br />

or bulldozing. So very little, if anything, is being done to control it.<br />

Herbicides give unsatisfactory results and 2,4,S-T which was recommended as a<br />

treatment <strong>of</strong> freshly cut stumps is now outlawed in many states and provinces (Hanson<br />

& Grau 1979). In any event, combinations <strong>of</strong> 2,4,-D and 2,4,S-T had little permanent<br />

effect and treated bushes regrew from the base and many <strong>of</strong> the test buckthorn leafed out<br />

one year after treatment (Switzer 1961). Ammate-X (ammonium sulfamate, duPont) is<br />

recommended as either a foliar spray or a stump treatment in Wisconsin (Hanson &<br />

Grau 1979) but early tests in Ontario showed that Ammate-X and fenuron (3-phenyl-l, 1dimethylurea)<br />

had little effect in that most <strong>of</strong> the test plants over four feet high continued<br />

to grow (Switzer et al. 1960, Switzer 1961).<br />

Cutting buckthorn gives at best temporary control because the stumps send up<br />

vigorous shoots in a relatively short time. However, cutting <strong>of</strong> individual bushes can<br />

reduce the incidence <strong>of</strong> crown rust in nearby fields.<br />

(e) Biological control<br />

Because chemical control <strong>of</strong> buckthorn is at best temporary, biological control is an<br />

attractive alternative and so at the request <strong>of</strong> <strong>Canada</strong> a survey for biological control<br />

agents was made by tbe Commonwealth Institute <strong>of</strong> Biological Control in 1964<br />

(Malicky et al. 1970). It was found that practically all the phytophagous insects specific<br />

to or closely associated with Rhamnaceae are in the orders Lepidoptera and<br />

Hemiptera. Many Coleoptera were found on Rhamnaceae but none was specific<br />

(Malicky el al. 1970),<br />

Investigations. particularly starvation tests suggest that the geometrids SCOlosia<br />

vetulata Schiff., Triphosa dubitata L., and the Iycaenid Thecla spini L. may be<br />

possible species for the biological control <strong>of</strong> R. cathartica in <strong>Canada</strong>. The geometrids<br />

combine a high degree <strong>of</strong> host specificity with a potential to defoliate their host,<br />

however. T. dubitata overwinters as an adult which may prove difficult in Canadian<br />

situations. T. spin; may have too wide a host range to be an acceptable agent, but<br />

there is some doubt as to the correctness <strong>of</strong> the host records (Malicky el al. 1970).<br />

R. calharlica is only one <strong>of</strong> several species <strong>of</strong> Rhamnaceae capable <strong>of</strong> being an<br />

alternate host to crown rust <strong>of</strong> oats (Dietz 1926). It is, however, the only one occurring<br />

in any significant numbers in <strong>Canada</strong> (Peturson 19S4) .. Eradication <strong>of</strong> buckthorn would<br />

not erase the problems <strong>of</strong> crown rust but would remove the site <strong>of</strong> the aecial or sexual<br />

spores <strong>of</strong> the rust and thus the source <strong>of</strong> pathogenic variation through hybridization<br />

and recombination <strong>of</strong> genetic material (Simons el al. 1979). Hybridization and recom-


188 M. G. Maw<br />

Recommendations<br />

Literature Cited<br />

bination, however, may not be as important as once thought. New races appear first in the<br />

main north-south rust population, and only later become established in populations<br />

cycling on buckthorn. (Simmons M.D. personal communication 1983).<br />

Infections <strong>of</strong> rust due to buckthorn are usually very local but can result in complete<br />

loss <strong>of</strong> the crop near buckthorn. Infection levels usually decline linearly with distance<br />

from the buckthorn (Harder 1975) and a few hundred meters from the source is<br />

sufficient to make a significant difference in infection incidence. Growers must either<br />

avoid planting oats near buckthorn or eradicate it from their farms. Buckthorn is so<br />

widespread and persistent that complete eradication is not feasible.<br />

Biological control <strong>of</strong> buckthorn appears to be the only course to lessen the impact<br />

<strong>of</strong> crown rust on oat production. It is not expected that introduced bi;:'logical agents<br />

will eradicate buckthorn, but they might add to the stress <strong>of</strong> climate and soil type to<br />

slow the spread <strong>of</strong> the shrub. Also, it is expected that any reduction in leaf surface will<br />

aid in diminishing rust inoculum early in the season when the greatest injury is caused<br />

to the oat crop. Not many insects find buckthorn an attractive food source because <strong>of</strong><br />

the feeding deterrent chemical, emodin (Trial & Dimond 1979). The original sources <strong>of</strong><br />

the North American populations <strong>of</strong> buckthorn are not known and there could be<br />

different biotypes containing varying amounts <strong>of</strong> emodin. This may make selection <strong>of</strong><br />

effective biological control agents difficult.<br />

Surveys for phytophagous insects in Europe found no species specific to R.<br />

eathartiea, although the Lepidoptera Seotosia vetulata and Thecla spini exhibit a<br />

degree <strong>of</strong> specificity to Rhamnaceae.<br />

Thus when attempting to control buckthorn biologically, one has to consider that<br />

species other than R. eathartiea will be attacked. The degree to which these other<br />

species will be damaged must be determined in the screening tests, and decisions made<br />

as to whether damage and possible loss <strong>of</strong> other species <strong>of</strong> Rhamnaceae can be<br />

tolerated.<br />

Even though R. eathartiea is the source <strong>of</strong> early infection <strong>of</strong> oats by crown rust and is the<br />

site <strong>of</strong> the aecial stage <strong>of</strong> the rust, its complete eradication would at best result in only a<br />

reduced incidence <strong>of</strong> the disease and the appearance <strong>of</strong> new rust races. Therefore. it is<br />

recommended that biological control work on R. cathartiea be discontinued.<br />

Bassett. 1.1. (1958) A sUlVey <strong>of</strong> the European buckthorn and common barberry in southern Manitoba. Science Service, Division <strong>of</strong> Botany<br />

and Plant Pathology, Canadian Department <strong>of</strong> Agriculture Ottawa, 22 pp.<br />

Dietz, S.M. (1926) The alternate hosts <strong>of</strong> crown rust, PucciniD cor<strong>of</strong>Ulta Corda. Jounuzl 0/ Agricultural Research 33, 953-970.<br />

Fernald. M.L. (1950) Gray's manual <strong>of</strong> botany. American Book Corp., New York, 1632 pp.<br />

fleischmann. G. (1963) Crown rust <strong>of</strong> oats in <strong>Canada</strong> in 1963. ConodiDn Plant Disease Survey 43,168-172.<br />

fleischmann, G. (1968) Crown rust <strong>of</strong> oats in <strong>Canada</strong> in 1968. CatuldiDn Plant Disease Survey 48,99·101.<br />

fleischmann, G. (1969) Crown rust <strong>of</strong> oats in <strong>Canada</strong> in 1969. Canadian Plant Disease Survey 49,91-94.<br />

fleischmann. G.; Samborski, D.l.; Peturson, B. (1963) The distribution and frequency <strong>of</strong> occurrence <strong>of</strong> physiologic races <strong>of</strong> Puccinia<br />

coronata Corda. f. sp. avenae Erikas. in <strong>Canada</strong>. CanaditJn Journal <strong>of</strong> Botany 41, 481-487.<br />

Godwin, H. (1943) Rhamnaceae. Journal <strong>of</strong> Ecology 31, 66-92.<br />

Hanson, E.W.; Gmu, G.R. (1979) The buckthorn menace to oal production. University <strong>of</strong> Wisconsin Extension Bulletin A 2860,2 pp.<br />

Harder, D.E. (1975) Crown rust <strong>of</strong> oats in <strong>Canada</strong> in 1974. Canaditln Plant Disease Survey 55, 63-65.<br />

Harder, D.E. (1978) Crown rust <strong>of</strong> oats in <strong>Canada</strong> in 19n. CafUldian Plant Disease Survey 58, 39-43.<br />

Harder, D.E.; McKenzie, R.l.H. (1974) Crown rust <strong>of</strong> oats in <strong>Canada</strong> in 1973. CafUldian Plant Disease Survey 54, 16-20.<br />

Malicky, H.; Sobhian, R.; Zwolfer, H. (1970) Investigations on the possibilities <strong>of</strong> a biological control <strong>of</strong> Rhamnus cathartica L. in<br />

<strong>Canada</strong>. Host mnges, feeding sites, and phenology <strong>of</strong> insects associated with European Rhamnaceae.<br />

Zriuchri/t fUr angewantlte Entomologie 6S, n -97.


Blank Page<br />

190


Pest Status<br />

Background<br />

Chapter 40<br />

Salsola pestiter A. Nels., Russian<br />

Thistle (Chenopodiaceae)<br />

P. HARRIS<br />

Russian thistle, Salsola peslifer A. Nels. (= S. iberica Sennen & Pau; = S. kali var.<br />

lenuifolia Tausch.) is an annual, native to southern Russia and western Siberia. It was<br />

introduced as a flax seed contaminant into South Dakota in 1873 (Robbins el al. 1951),<br />

but it was probably also present in the many introductions <strong>of</strong> Turkestan alfalfa imported<br />

into <strong>Canada</strong> and the United States. The reason for suspecting this source was that Wilcox<br />

& Stevenson (1909) found it in 51 out <strong>of</strong> 201 samples <strong>of</strong> alfalfa seed in Nebraska. In<br />

<strong>Canada</strong>, it became particularly prevalent in the 19305: Moynan (1939) reported Russian<br />

thistle constantly invading new territory in Manitoba, and that between 1935-38 it<br />

became established on illustration stations at Roblin, Gilbert Plains, and Ste. Rose. It is<br />

now present in all provinces except Newfoundland, although in eastern <strong>Canada</strong> it is<br />

almost exclusively a railway and roadside weed (Frankton & Mulligan 1970). It is,<br />

however, abundant in the drier parts <strong>of</strong> the prairies and British Columbia on both<br />

cultivated land and over-grazed or disturbed pasture. In Saskatchewan, Russian thistle<br />

varied between the fifth to the seventh most abundant weed on cultivated land and in the<br />

southwest <strong>of</strong> the province it was present in 64% to 100% <strong>of</strong> the 0.25 ml samples surveyed<br />

(Thomas 1976,1977,1978, 1979). In winter the plants become tumble weeds that collect<br />

on fences and other obstructions.<br />

In western United States, Russian thistle is the favored alternate host <strong>of</strong> the beet<br />

leafhopper, Circulifer lenellus (Baker), which is the vector <strong>of</strong> the curly top virus <strong>of</strong> sugar<br />

beet and other crops (Severin 1933). The losses caused by this disease have been a major<br />

impetus for the biological control <strong>of</strong>the weed in the United States, however, the disease<br />

does not occur in the sugar beet areas <strong>of</strong> either Manitoba (Robertson 1968) or Alberta<br />

(R.J. Howard, 1981, personal communication).<br />

The insect guild on Russian thistle in North America is typical <strong>of</strong> an introduced weed.<br />

Goeden & Ricker (1968) found 91 species <strong>of</strong> insects feeding on it, but none <strong>of</strong> them did<br />

appreciable damage and all but two were polyphagous. In contrast, in the palaearctic the<br />

plant is attacked by a number <strong>of</strong> specialized insects (Hawkes 1969, Goeden 1973). Two<br />

<strong>of</strong> these have been cleared by workers in the United States. The Canadian programme<br />

against Russian thistle has been based on this work.<br />

The stem boring moth, Coleophora parthenica Meyr., was released in the United<br />

States in 1973 following host specificity studies by Hawkes & Mayfield (1976). The<br />

species has a high temperature threshold: a minimum <strong>of</strong> IS.5°C is required for development<br />

and about 500 degree-days above this temperature for maturation; however, about<br />

twice this number <strong>of</strong> degree-days are required for a population increase (Hawkes, 1976,<br />

personal communication). The moth is well established and has attained high population<br />

densities at many places in southwestern United States where it normally has three<br />

generations a year (Hawkes & Mayfield 1978b). At Indio, California, Goeden & Ricker<br />

(1979) found an average <strong>of</strong> one larva per 7.4 cm <strong>of</strong> stem. This level <strong>of</strong> attack had no<br />

effect on plant size, mortality. or seed production so the moth will have to be supplemented<br />

with other biological control agents to achieve beneficial results.<br />

The case bearer. Coleophora klimeschiella Toll., was released in California in 1977<br />

191


192 P. Harris<br />

Releases and Recoveries<br />

Table 49<br />

Evaluation <strong>of</strong> Control Attempts<br />

following screening studies by Hawkes & Mayfield (19780). The biology <strong>of</strong> this insect<br />

is reported by Khan & Baloch (1976): the first two instars are leafminers; in the last<br />

three, the larva lives outside the leaf in a case <strong>of</strong> hollowed leaves, and feeds by<br />

protruding its head and forelegs through a hole in the leaf so that it continues to feed<br />

inside the leaf. The insect overwinters as a larva inside the case attached to SaJsoJa<br />

stems or stones. There are one to three generations a year in Pakistan, depending on<br />

elevation. C. klimeschiella is established in California (Julien 1982).<br />

The releases <strong>of</strong> both C. parthenica and C. klimeschielJa are listed in Table 49.<br />

C. parthenica was released on a number <strong>of</strong> occasions in July and August in stands <strong>of</strong><br />

Russian thistle at the Regina and Swift Current Research Stations. The moths remained<br />

on the plants where they were released for several days, but no eggs or larvae were<br />

found in the vicinity. Larvae in the stems <strong>of</strong> potted plants that were placed outside failed<br />

to complete development. The problem was presumed to be the cool summer temperatures<br />

<strong>of</strong> the Canadian prairies. At Regina in July and August, the minimum temperature<br />

was below the threshold for development for 50 days in July and August, and on 16<br />

August, the temperature fell to 1.3°C. At all sites on the Canadian prairies, there is<br />

normally less than the 500 degree-days required by the insect.<br />

C. klimeschielJa was released in June at the Regina Research Station and bred in the<br />

summer <strong>of</strong> release and the following summer. The population was less than one larva per<br />

Russian thistle plant and they did no appreciable damage. In the spring <strong>of</strong> 1979, the<br />

release site was flooded for several weeks in the spring and the colony was destroyed.<br />

Open releases and recoveries against SaJsola pestifer A. Nels.<br />

Year No. <strong>of</strong> moths released Source Release site<br />

Coleophora parthenica<br />

1975 273<br />

1975 66<br />

1975 33<br />

Coleophora klimeschiella<br />

1977 91<br />

California<br />

California<br />

Pakistan<br />

Swift Current. Saskatchewan<br />

Regina. Saskatchewan<br />

Regina. Saskatchewan<br />

California Regina, Saskatchewan<br />

ex Pakistan<br />

Recovery<br />

bred in 1977<br />

and 1978 but<br />

not in 1979<br />

Summer temperatures on the Canadian prairies are too low for C. parthenica, and there<br />

is no purpose in making additional releases unless a strain with a low temperature<br />

threshold can be found. Even if such a strain can be established in <strong>Canada</strong>, on the basis <strong>of</strong><br />

results in the United States, the density <strong>of</strong> Russian thistle is unlikely to be affected.<br />

C. klimeschiella can survive in Saskatchewan and hence presumably elsewhere in<br />

southern <strong>Canada</strong>. Insects released for the biological control <strong>of</strong> a weed <strong>of</strong>ten require<br />

several years for the selection <strong>of</strong> a strain adapted to the region, so the lack <strong>of</strong> a rapid<br />

increase in the population should not be the sole reason for discontinuing further


Recommendations<br />

Acknowledgements<br />

Literature Cited<br />

Sa/.m/a pestilL'r A. Nels.. 193<br />

attempts. More discouraging is the absence <strong>of</strong> any data on the losses caused by this<br />

extremely common weed on the prairies and the present general lack <strong>of</strong> interest in a<br />

biological control solution for Russian thistle.<br />

Biological control <strong>of</strong> S. pestifer should be discontinued until its use can be justified in<br />

economic terms.<br />

I am grateful to R.B. Hawkes for reviewing the manuscript.<br />

Frankton. C.: Mulligan. G.A. (1970) Weeds <strong>of</strong> <strong>Canada</strong>. Canadian Department <strong>of</strong> Agriculture. 217 pp.<br />

Goeden. R.D. (1973) Phytophagous insects found on Salsola in Turkey during exploration for biological weed control agents for California.<br />

Entomophaga 18. 439-448.<br />

Goeden. R.D.; Ricker. D.W. (1968) The phytophagous insect fauna <strong>of</strong> Russian thistle (Salsola kali var. tenuifolia) in southern California.<br />

Annals <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> America 61. 67-72.<br />

Goeden. R.D.; Ricker. D.W. (1979) Field analysis <strong>of</strong> Coleophora parthenica (Lcp.: Coleophoridae) as an imported natural enemy <strong>of</strong><br />

Russian thistle. Salsola iberica. in the Coachella valley <strong>of</strong> southern California. Environmental<br />

Entomology 8. 1099-1101.<br />

Hawkes. R.B. (1969) Insects recorded from Salsola spp. and Halogeton spp. Report <strong>of</strong> the Entomology Research Division. USDA. ARS.<br />

Albany. California. 55 pp.<br />

Hawkcs. R.B.; Mayfield. A. (1976) Host specificity and biological studies <strong>of</strong> Coleophora parthenica Meyrick. an insect for the biological<br />

control <strong>of</strong> Russian thistle. Commemorative Volume in Entomology. Department <strong>of</strong> Entomology.<br />

University <strong>of</strong> Idaho. 37-43.<br />

Hawkes. R.B.; Mayfield. A. (19780) Coleophora klimeschitlla. biological control agent for Russian thistle: host specificity testing.<br />

Environmental Entomology 7. 257 - 261.<br />

Hawkes. R.B.; Mayfield. A. (l978b) Coleophora spp. as biological control agents against Russian thistle. Proceedings <strong>of</strong> the 4th Inter.<br />

national Symposium on Biological Control <strong>of</strong> Weeds. 113-116.<br />

Julien. M.H. (1982) Biological control <strong>of</strong> weeds: a world catalogue <strong>of</strong> agents and their Iarget weeds. Commonwealth Institute <strong>of</strong> Biological<br />

Control. lOS pp.<br />

Khan. A.G.; Baloch. G.M. (1976) Coleophora klimeschiella (Lcp: Coleophoridae) a promising biocontrol agent for Russian thistles.<br />

Salsola spp. Entomophaga 21. 425-428.<br />

Moynan. J.e. (1939) Di\ision <strong>of</strong> illustration stations. Progress report 1934 to 1938. Part II. Department <strong>of</strong> Agriculture. <strong>Canada</strong>. 130 pp.<br />

Robertson. H. (1968) Sugar farmers <strong>of</strong> Manitoba. The Manitoba Beet Growers Association. 176 pp.<br />

Robbins. W.W.; Bellue. M.K.; Ball. W.S. (1951) Weeds <strong>of</strong> California. Sacramento; California State Printing Office. 547 pp.<br />

Severin. H.H.P. (1933) Field observations on the beet leafhopper. EUlellix ttnnt'llus. in California. Hilgardia 7. 281-360.<br />

Thomas. A.G. (1976) Weed survey <strong>of</strong> cultivated land in Saskatchewan. First annual report. Agricultural <strong>Canada</strong>. 93 pp.<br />

Thomas. A.G. (1977) Weed survey <strong>of</strong> cultivated land in Saskatchewan. Second annual report. Agriculture <strong>Canada</strong>. 103 pp.<br />

Thomas. A.G. (1978) Weed survey <strong>of</strong> cultivated land in Saskatchewan. Third annual report. Agriculture <strong>Canada</strong>. 113 pp.<br />

Thomas. A.G. (1979) Weed survey <strong>of</strong> cultivated land in Saskatchewan. Fourth annual report. Agriculture <strong>Canada</strong>. 141 pp.<br />

Wilcox. E.M.; Stevenson. N. (1909) Report <strong>of</strong> the Nebraska seed laboratory. Bulletin <strong>of</strong> the Agric-u1tural Experiment Station, Nebraska 110.<br />

29pp.


Blank Page<br />

194


Pest Status<br />

Background<br />

Chapter 41<br />

Senecio jacobaea L., Tansy Ragwort<br />

(Compositae)<br />

P. HARRIS, A.T.S. WILKINSON and J.H. MYERS<br />

The status <strong>of</strong> tansy ragwort, Senecio jacobaea L., in <strong>Canada</strong> was described by Harris el<br />

al. (1971). Since that time the distribution has changed slightly. A small stand <strong>of</strong> the<br />

weed that had existed at Guelph, Ontario, for many years recently started to spread (J.<br />

Alex, 1981, personal communication). A small infestation in the Gaspe Peninsula,<br />

Quebec, recorded prior to 1900 could not be found by Watson & Muirhead (1979). The<br />

weed has continued to spread in the lower Fraser Valley in British Columbia, and it<br />

remains a serious socio-economic problem in the maritimes where there are many small<br />

mixed farms with a few cattle. The small field size and lack <strong>of</strong> specialized equipment<br />

makes chemical and mechanical control difficult and expensive. On the other hand, the<br />

death <strong>of</strong> one or two cattle from ragwort poisoning is a serious loss to the family food and<br />

income. There are no provincially compiled statistics <strong>of</strong> cattle deaths from ragwort as<br />

diagnosis is difficult and many cattle are sent for early slaughter if ragwort poisoning is<br />

suspected. However, in 1981, in Prince Edward Island the Montague Veterinary Clinic<br />

reported 12-20 ragwort-related cattle deaths; the Kensington Clinic had 20 but many<br />

more were suspected; the Provincial Veterinary Pathology Laboratory reported 10-12<br />

but suspected more (L. Thompson, 1981, personal communication). Thus the total<br />

cattle mortality from ragwort is similar to the 40-65 deaths estimated to have occurred<br />

in 1968 (Harris et al. 1971).<br />

The cinnabar moth, Tyria jacobaeae (L.), was established in the Canadian maritimes in<br />

1964 and in British Columbia in 1965 for the biological control <strong>of</strong> tansy ragwort. Despite<br />

difficulties in getting the moth established (Harris et al. 1975), it is now present throughout<br />

the ragwort infested regions <strong>of</strong> <strong>Canada</strong>, including Newfoundland where no releases<br />

were made (Larson & Jackson 1980).<br />

The effect <strong>of</strong> the moth on the weed has varied widely. At Prince Charles, Prince<br />

Edward Island, and Durham, Nova Scotia, the weed has practically disappeared from<br />

the release sites although it has persisted in the disturbed ground along the roadsides<br />

(Harris el al. 1978). At Durham it was largely replaced by goldenrod Solidago sp. At<br />

Sussex, New Brunswick, the ragwort stand collapsed after defoliation by cinnabar<br />

larvae while the <strong>Canada</strong> thistle, Cirsium arvense (L.) Scop., increased. Tansy ragwort<br />

has since returned to the site and the density <strong>of</strong> cinnabar reached circa three larvae/stem<br />

in 1980 with scattered but numerous larvae in 1981 (D. Finnamore, 1981, personal<br />

communication). Near Nanaimo, British Columbia, over the past ten years an average<br />

<strong>of</strong> 38% <strong>of</strong> the ragwort was defoliated with another 18% partially defoliated. This has had<br />

little effect on the density <strong>of</strong> the flowering plants but they are smaller. In California,<br />

Hawkes & Johnson (1978) found that, with cinnabar moth attack, the number and size<br />

<strong>of</strong> the ragwort flowering stems decreased and the number <strong>of</strong> rosettes increased. Not all<br />

the declines were the result <strong>of</strong> the cinnabar moth as Myers (1980) found that two <strong>of</strong> five<br />

stands in Nova Scotia declined even though the cinnabar population was not large<br />

enough to cause extensive defoliation.<br />

Many papers have been written in <strong>Canada</strong> and elsewhere on the causes <strong>of</strong> the diverse<br />

effects <strong>of</strong> the cinnabar moth on tansy ragwort. Harris et al. (1978a) related the collapse<br />

195


196 P. Harris. A. T. S. Wilkinson and J. H. Myers<br />

<strong>of</strong> some maritime stands <strong>of</strong> the weed following cinnabar defoliation to the frost sensitivity<br />

<strong>of</strong> the regenerating plants. Several studies (Green 1974, Campbell 1975, Myers 1976,<br />

Myers & Campbell 1976a, 1976b) have investigated the importance <strong>of</strong> ragwort spacing<br />

on cinnabar egg distribution and the success <strong>of</strong> larval transfer between plants. They<br />

showed that one requirement for explosive increases <strong>of</strong> the moth was dense stands <strong>of</strong> the<br />

weed. Dempster (1971) reported a higher mortality <strong>of</strong> cinnabar eggs and larvae on<br />

ragwort rosettes than on flowering stems and Myers (1980) found that cinnabar populations<br />

in areas with a high density <strong>of</strong> closely spaced rosettes tended to be less stable than<br />

on those with low rosette density. A high nitrogen content in the plant increased both<br />

moth fecundity and egg mass size in the following generation. Myers & Post (1981)<br />

argued that this would tend to destabilize the cinnabar population by over-exploitation<br />

<strong>of</strong> the food resource. In a study by Lakhani & Dempster (1981), the number <strong>of</strong> eggs had<br />

a rather small effect on the subsequent moth numbers because high egg density was<br />

<strong>of</strong>ten associated with larval starvation. In a dune habitat, Meijden (1979) described the<br />

cinnabar population as "walking on ice floes" because small stands <strong>of</strong> ragwort became<br />

extinct one or two years after attack by the moth and then reappeared after the moth<br />

emigrated to other stands. Predators were important in some instances (Wilkinson 1965,<br />

Myers & Campbell 1976c, Meijden 1979). In 1980 at Nanaimo, British Columbia, most <strong>of</strong><br />

the pupae were destroyed resulting in a low population in 1981 in which only 3% <strong>of</strong> the<br />

plants were attacked.<br />

Philogene (1975) found that the day length and temperature during larval development<br />

did not affect the obligatory pupal diapause. However, the moth has adapted its temperature<br />

threshold for emergence in the spring so that in the various regions <strong>of</strong> North<br />

America larval feeding remains synchronized with ragwort flowering (Myers 1979).<br />

Richards & Myers (1980) found that maternal moth size and temperature requirements<br />

for moth emergence were heritable and Myers (1978) found that the present populations<br />

<strong>of</strong> the moth have regional differences in their enzyme systems that are irrespective <strong>of</strong><br />

their origin. Thus in a few years since release, the moth has adapted to its new habitat<br />

with the result that its behaviour now is not necessarily the same as that <strong>of</strong> the stock<br />

released.<br />

A model by Lakhani & Dempster (1981) showed that for both Nanaimo, British<br />

Columbia, and Weeting Heath in Britain, the changes in the density <strong>of</strong> the weed closely<br />

conformed to the amount <strong>of</strong> spring and early summer rainfall. The model showed that at<br />

these sites the cinnabar population merely tracked the changes in the density <strong>of</strong> the<br />

weed. Myers (1980) also concluded that population fluctuations after an initial reduction<br />

<strong>of</strong> plant size by introduced cinnabar moths have a large environmental component. One<br />

indication that different factors are important in different sites is that the model did not<br />

fit the results from a more moist site in Oregon. In a more generalized model R<strong>of</strong>f &<br />

Myers (unpublished) found that depending on the degree <strong>of</strong> larval dispersal, the cinnabar<br />

population could be cyclic, stable, or chaotic. The gamut <strong>of</strong> these conditions<br />

occurs in nature. There are several ragwort habitats that are not utilized by the moth: the<br />

weed tends to be avoided when growing in partial shade; the cinnabar pupae are not able<br />

to overwinter in wet sites (Dempster 1971) so the moth tends to be absent in Europe from<br />

pastures on river flood plains, which <strong>of</strong>ten have dense ragwort stands; the moth is rare in<br />

the Swiss Jura on the widely separated plants or small clumps <strong>of</strong> plants. It is a weak flier<br />

and has dispersed less rapidly in California than the beetle Longitarsus jacobaeae<br />

(Waterhouse).<br />

In many habitats the cinnabar moth has not reduced the density <strong>of</strong> tansy ragwort, but<br />

plants defoliated annually tend to be smaller so there has been some reduction in ragwort<br />

biomass. Also for about two months in the summer there is little ragwort foliage in the<br />

pastures or hay fields so that availability <strong>of</strong> the toxic foliage to cattle is reduced. The<br />

level <strong>of</strong> control is not satisfactory as a high density <strong>of</strong> the weed remains on many sites for<br />

much <strong>of</strong> the year, so the cinnabar moth needs to be supplemented by additional agents.


Releases and Recoveries<br />

HyJemya senecieJJa<br />

Meade (Diptera:<br />

Muscidae)<br />

Tyriajacobaeae (L.)<br />

(Lepidoptera:<br />

Arctiidae)<br />

(a) Ecology<br />

Senecio j(lcoh(l('(l L. . 197<br />

H. senecie/la larvae reduce seed production by feeding in ragwort flower heads. The<br />

biology was described by Miller (1970). and an unsuccessful attempt to establish it in<br />

British Columbia and Prince Edward Island in 1968 was reported by Harris el 01. (1971).<br />

The fly has been established in New Zealand where it infested up to 77% <strong>of</strong> the heads at<br />

peak flowering but the effect <strong>of</strong> the reduced seed production on ragwort density was not<br />

determined. The fly was established in California and probably in Oregon and<br />

Washington (Frick 19690) but the California site was subsequently destroyed (Julien<br />

1981).<br />

(b) Releases<br />

A summary <strong>of</strong> releases in the current review period is shown in Table 50; none <strong>of</strong> them<br />

became established. At least part <strong>of</strong> the problem was poor synchrony between emergence<br />

<strong>of</strong> the flies and ragwort flowering: in 1970 the H. seneciella were received in May<br />

and placed outside in a field cage in Prince Edward Island. The flies emerged between 8<br />

and 17 June but the host plant did not flower until August. In 1971 emergence was<br />

delayed until the second half <strong>of</strong> July but still did not result in establishment (L.<br />

Thompson. 1981. personal communication). The results from British Columbia were<br />

similar.<br />

Establishment in the United States was obtained with much larger releases <strong>of</strong> circa<br />

2000 flies. As egg viability was circa 25% (Frick 1969b) a larger release than that used in<br />

<strong>Canada</strong> is necessary. The problem with the synchronism <strong>of</strong> fly emergence can probably<br />

be overcome by releasing the mature larvae at the end <strong>of</strong> the summer; this could be done<br />

with the Oregon or Washington stock providing that it is free <strong>of</strong> parasites. The value <strong>of</strong><br />

establishing the fly is less clear since most ragwort reproduction in pastures is by vegetative<br />

propagation from existing plants. It is unlikely that the fly would survive in stands<br />

defoliated by cinnabar larvae which consume the flowers preferentially.<br />

Releases<br />

Field collected larvae from Nova Scotia were released at two sites in New Brunswick in<br />

1970 and at seven sites in Ontario in 1979 and 1981 (Table 50). These were the only<br />

releases since those reported by Harris el 01. (1971). Establishment occurred at both<br />

sites in New Brunswick resulting in widespread defoliation <strong>of</strong> tansy ragwort but on a<br />

long-term basis this is a less than desirable level <strong>of</strong> control. The insect did not become<br />

established in Ontario, possibly because <strong>of</strong> predation by ants (Alex, 1981, personal<br />

communication). About 20% <strong>of</strong> the Ontario shipment died in transit indicating that the<br />

popUlation was heavily stressed and probably diseased.


Recommendations<br />

Table 50<br />

Senecio jacobaI'll L. , 199<br />

L. jacobaeae was recovered at the two Prince Edward Island sites in 1982 and 1983<br />

although numbers were low (L.S. Thompson personal communication 1983). Also the<br />

possibility <strong>of</strong> some root damage from larval feeding was noted in New Brunswick (D.<br />

Finnamore personal communication 1981). Prince Edward Island has a more even<br />

distribution <strong>of</strong> summer rain than occurs on the south west coast <strong>of</strong> British Columbia, so<br />

possibly the summers are too moist for the biotype released. Frick & Johnson (1972)<br />

found that the life cycle <strong>of</strong> the Swiss biotype adapted in the laboratory by a decrease in the<br />

length <strong>of</strong> the egg diapause. With the beetle's rapid adaptability and the probable wide<br />

genetic variation <strong>of</strong> the British Columbia population (stock from three sources has been<br />

released there), it should be possible to select an adapted biotype in a field cage and<br />

increase the survivors in the laboratory if the field population falls critically low. The<br />

other alternative is to import beetles from a climate with a similar summer rainfall to<br />

Prince Edward Island, but even so some adaptation would probably be necessary.<br />

(1) The cinnabar moth has reduced the seriousness <strong>of</strong> tansy ragwort in many areas;<br />

however additional species <strong>of</strong> biological control agents are needed if a fully satisfactory<br />

level <strong>of</strong> ragwort control is to be achieved.<br />

(2) The most obvious candidate as a second biological control agent is the beetle<br />

Longilarsils jacobaeae. A biotype <strong>of</strong> this beetle is now adapted to the southwest coast <strong>of</strong><br />

British Columbia and should be distributed to ragwort infestations in the region. Estab-<br />

Open releases and recoveries against Senecio jacobaea L.<br />

Species and release site Year Origin No. Recovered<br />

Long;tarsus jacobaeae (Waterhouse)<br />

British Ollumbia<br />

Vancouver 1971 California ex Italy 100 adults 1974<br />

Vancouver 1973 Switzerland 38 adults<br />

Abbotsford 1971 California ex Italy 400 adults<br />

1972 Lab reared ex California<br />

and England 400 adults<br />

1972 California and England 250 adults<br />

1974 UBC site ex Switzerland 200 adults 1975<br />

Nanaimo 1974 England 92 adults 1976<br />

1976 Oregon ex Italy 1000 adults<br />

Chilliwack 1978 Abbotsford 500 adults 1979<br />

Prince Edward Island<br />

Mt. Herbert 1978 British Columbia 300 adults 1982<br />

Charlottetown 1981 British Columbia ISS adults 1982<br />

New Brunswick<br />

Fredericton 1981 British Columbia 180 adults<br />

Tyrill jacobaeat (L.)<br />

New Brunswick<br />

Sussex 1970 Nova Scotia 1000 larvae 1971<br />

Bathurst 1970 Nova Scotia 1000 larvae 1971<br />

Ontario<br />

Guelph 1979 Nova Scotia 4500 larvae Not recovered<br />

Guelph 1981 Nova Scotia 4500 larvae Not recovered<br />

H)'lemya sm«iella (Meade)<br />

British Ollumbia<br />

Abbotsford 1970 Switzerland 6S adults<br />

453 puparia Not recovered<br />

Prince Edward Island<br />

Charlottetown 1970 Switzerland liS adults Not recovered<br />

Charlottetown 1971 United States liO adults Not recovered


Sellecio jacoba('(/ 201<br />

Myers, 1.H. (1980) Is the insect or the plant the driving force in the cinnabar moth-tansy ragwort system? Oecologia 47.16-21.<br />

Myers. J.H.: Campbell. B.l. (19700) Indirect measures <strong>of</strong> larval dispersal in the cinnabar moth. Tyriajacobal'al' (Lepidoptera: Arctiidae).<br />

Canadian Entomologi5/ lOB. 967-972.<br />

Myers. 1.H.: Campbell. B.l. (1976b) Distribution and dispersal in populations capable <strong>of</strong> resource depletion. A field study on cinnabar<br />

moths. Oecologia 24. 7-20.<br />

Myers, J.H.: Campbell, B.l. (1976c) Predation by carpenter ants: a deterrent to the spread <strong>of</strong> cinnabar moth. Journal <strong>of</strong> the <strong>Entomological</strong><br />

<strong>Society</strong> <strong>of</strong> British Columbia 73. 7-9.<br />

Myers, I.H.: Post. B.I. (1981) Plant nitrogen and fluctuations <strong>of</strong> inM:ct populations: a test with the cinnabar moth-tansy ragwon system.<br />

Oemlogia 48, 151-156.<br />

Newton. H.C.F. (1933) On the biology <strong>of</strong> some species <strong>of</strong> Longi;ar.sus living on ragwon. Bulletin <strong>of</strong> <strong>Entomological</strong> Research 24. 511-520.<br />

Philogene. B.J.R. (1975) Responses <strong>of</strong> the cinnabar moth, Hypocrita jacobaeae to various temperature/photoperiod regimes. Journal <strong>of</strong><br />

Insect Physiology 21. 1415-1417.<br />

Richards. L.J.: Myers, 1.H. (1980) Maternal influences on size and emergence time <strong>of</strong> the cinnabar moth. Canadian Journal <strong>of</strong> Zoology 58.<br />

1452-1457.<br />

R<strong>of</strong>f, D.; Myers. J.H. Simulating the dynamics <strong>of</strong> the cinnabar moth populations: a study <strong>of</strong> complexity in a deterministic world<br />

(unpublished manuscript).<br />

Shute, S.L. (1975) Longitarsw jacobaeae Waterhouse (Col.. Chrysomelidac): identity and distribution. Entomologists' Monthly<br />

Magazine. 111 (1328-13300). 33-39.<br />

Watson, A.K.; Muirhead. L. (1979) Exploration <strong>of</strong> the Gaspe region <strong>of</strong> Quebec for tansy ragwon. p. 41 I. In: Research repon. Expen<br />

Committee on Weeds (Eastern <strong>Canada</strong>). 412 pp.<br />

Wilkinson. A.T.S. (19(5) Releases <strong>of</strong> cinnabar moth. Hypocrita jacobaeae (L.) (Lepidoptera: Arctiidae) on tansy ragwon in British<br />

Columbia. Proceedings <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> British Columbia 62. 10-13.


Blank Page<br />

202


Pest Status<br />

Background<br />

Chapter 42<br />

Silene cucubalus Wibel, Bladder Campion<br />

(Caryophyllaceae)<br />

M.O. MAW<br />

Silene cucubalus Wibel is a deep rooted perennial <strong>of</strong> Eurasian origin. It is spread by seed<br />

(as many as 20 000 per plant) as well as vegetatively by parts <strong>of</strong> the root crown severed<br />

by implements. It is resistant to the common herbicides 2,4-D (2,4-dichlorophenoxyacetic<br />

acid) and MCPA (4-chloro-2-methylphenoxyacetic acid) at rates which can be<br />

safely used in grain crops. The weed grows in waste places, roadsides, and railyards,<br />

and although it is an infrequent weed in cultivated ground (Boivin 1968), it can be<br />

troublesome in well drained pastures, cereal crops, and hay. It is found in the northeastern<br />

and central United States and in every province <strong>of</strong> <strong>Canada</strong> to latitude 54°N<br />

(Scoggan 1978) and is more common in the eastern than in the western parts <strong>of</strong> its range<br />

(Frankton et al. 1970).<br />

Although the weed is generally not a great problem throughout its Canadian range,<br />

there are localized areas where it causes concern. For example, in the Ethelbert district<br />

<strong>of</strong> Manitoba, a survey <strong>of</strong> weeds in cultivated fields (Thomas 1978) showed it to be<br />

present in 25% <strong>of</strong> the samples in 50% <strong>of</strong> the fields surveyed with a mean density <strong>of</strong> 5.9<br />

plants per square meter. In contrast, the weed was reported in only one <strong>of</strong> the remaining<br />

38 Manitoba districts surveyed. and there it was found only in 0.4% <strong>of</strong> the samples in<br />

8.3% <strong>of</strong> the fields at a density <strong>of</strong> 0.6 plants per square meter.<br />

Although it would seem that there has been little change in the status <strong>of</strong> the weed on<br />

cultivated lands since the 1966 survey (Alex 1966), agricultural practices have altered<br />

the situation in areas such as in southeastern Manitoba. Here a shift has been away from<br />

annual tillage to establishment <strong>of</strong> pastures and forage crops and forage seed production.<br />

Where bladder campion was not a problem as recently as three years ago, it now is a<br />

concern in over 4200 hectares and only limited control is being provided by herbicides<br />

and cultural methods.<br />

Surveys <strong>of</strong> the European insect fauna on species <strong>of</strong> Silene and Melandriwn were made<br />

by the Commonwealth Institute <strong>of</strong> Biological Control (Miotk 1973). Sixty-six percent <strong>of</strong><br />

the phytophagous insects on the two genera were oligophagous and the supposedly<br />

stenophagous species were either weevils or noctuids.<br />

The chrysomelid, Cassida azuna Fab. (mistakenly identified as Cassida hemisplwerica<br />

Hbst.) (Maw & Steinhausen 19800. 1980b), was collected in Switzerland and screened in<br />

<strong>Canada</strong> (Maw 1976). The Cassidinae are generally specialized in their feeding habits and<br />

species in a subfamily are usually restricted to a limited group <strong>of</strong> plants. This was found<br />

to be so with C. azurea.<br />

First instar larvae fed on all the plants in the tribes Sileneae and Alsineae tested but<br />

they developed only on S. cucubalus. S. cserei. S. glauca. S. noctiflora. S. maritima, S.<br />

alba, S. acau/is, Gypsophila repens, Dianthus chinensis .. and D. plumarius. Damage by<br />

young larvae is confined to the epidermis <strong>of</strong> the leaf. while older larvae and adults eat<br />

large holes in the leaf blade. Flowers <strong>of</strong> Silene are eaten by both larvae and adults.<br />

203


2M M. G. Maw<br />

Discussion<br />

Recommendations<br />

Literature Cited<br />

Although starvation tests give the impression that C. azurea has a wide host range, only<br />

plants in the family Caryophyllaceae supported development and in most instances only<br />

the younger tender leaves were eaten. Older and harder leaves and those with waxy<br />

coatings resist attack from young larvae and older larvae fare little better. It is expected<br />

that under field conditions, C. azurea would direct its attack to S. cucubalus. Other<br />

Silene would at best be marginal hosts. Dianthus spp. might be eaten when Silene spp.<br />

are unavailable. This would, however, be very restricted because the hard texture and<br />

waxy coating <strong>of</strong> the leaves make sustained feeding difficult for older larvae and adults<br />

and almost impossible for young larvae.<br />

With the main concern being the spread <strong>of</strong> S. cucubalus into forage, hay, and forage<br />

seed production fields, leaf feeding biological control agents may not be too successful.<br />

They will however be useful in situations where bladder campion is not cut, such as in<br />

waste areas. Root feeders should be <strong>of</strong> more use and so should be actively sought for<br />

biological control <strong>of</strong> S. cucubalus in hay and forage crops.<br />

(1) A biological control programme for S. cucubalus be continued.<br />

(2) C. azurea be reconsidered as a control agent.<br />

(3) Root feeders such as Hadena luteago Schiff. and H. andalusica Stgr. be<br />

studied and their potential as control agents be assessed.<br />

Alex, J .F. (1966) Survey <strong>of</strong> weeds <strong>of</strong> cultivated land in the prairie provinces. Regina, Saskatchewan; <strong>Canada</strong> Agriculture, Research Branch,<br />

68 pp.<br />

Boivin, B. (1968) Flora <strong>of</strong> the prairie provinces. Pan 2. Provancherea 3. Herbier Louis-Marie. Quebec; Universite Laval. 185 pp.<br />

Frankton. C.; Mulligan. G.A.; Wright, W.H.; Steins. I. (1970) Weeds <strong>of</strong> <strong>Canada</strong>. Ottawa; Canadian Depanment <strong>of</strong> Agriculture. 217 pp.<br />

Maw. M.G. (1976) Biology <strong>of</strong> the tonoise beetle Cossilill hemisphaerica (Coleoptera: Chrysomelidae), a possible biological control agent<br />

for bladder campion. Silene cucuba/us (Caryophyllaceac), in <strong>Canada</strong>. Canadian Entomologist 108.<br />

945-954.<br />

Maw. M.G.; Steinhausen. W.R. (19800) Corrigendum for "Biology <strong>of</strong> the tonoise beetle Cassilill hemisphaerica (Coleoptera: Chrysomelidae),<br />

a possible biological control agent for bladder campion Silene cucubalus (Caryophyllaceae). in <strong>Canada</strong>.<br />

Canadian Entomologist 112. 639.<br />

Maw. M.G.; Steinhausen, W.R. (1980b) Cassida azurta (Coleoptera: Chrysomelidae) - not C. hemisphaerica - as a possible biological<br />

control agent <strong>of</strong> bladder campion. Silene cucubalus (Caryopbyllaceae) in <strong>Canada</strong>. Ztitschri{t fur<br />

ange",andte Entomologie 90.420-422.<br />

Miotk. P. (1973) Phytophagous insects associated with weeds in central Europe. Pan 11. Wced Projects for <strong>Canada</strong>. Progress Repon 32.<br />

Commonwealth Institute <strong>of</strong> Biological Control. Delc!mont. Switzerland, 29 pp.<br />

Scoggan, H.J. (1978) The flora <strong>of</strong> <strong>Canada</strong>. Pan 3. Ottawa; National Museum <strong>of</strong> Natural Sciences. pp. 547-1115.<br />

Thomas. A.G. (1978) The 1978 weed survey <strong>of</strong> cultivatcd land in Manitoba. Regina; Agriculture <strong>Canada</strong> Research Station, 109 pp.


Pest Status<br />

Chapter 43<br />

Sonchus arvensis L., Perennial Sowthistle,<br />

S. oleraceus L., Annual Sowthistle<br />

and S. asper (L.) Hill, Spiny Annual<br />

Sow-thistle (Compositae)<br />

D.P. PESCH KEN<br />

Four species in the genus Sonchlls have been introduced into North America from<br />

Europe and Asia. Three <strong>of</strong> them are weeds. Perennial sow-thistle. Sonchus arvensis L..<br />

is a common weed in all provinces <strong>of</strong> <strong>Canada</strong> but it is particularly abundant in Ontario.<br />

Manitoba. and the northern agricultural areas <strong>of</strong> Quebec. Ontario. and the Prairie Provinces<br />

(Frankton & Mulligan 1970). It includes the varieties S. arvensis L. s. str. and S.<br />

arvensis var. glabrescens Guenth .• Grab & Wimm. (Boulos 1961). It reproduces by<br />

seed and an extensive underground root system. About 9750 achenes are produced per<br />

flowering shoot (Stevens 1932) and 59500 per m 1 (Stevens 1924). The seeds are dispersed<br />

by means <strong>of</strong> a pappus. Stevens (1924) noted that most seeds are not carried far<br />

and that there is a difference in the persistence <strong>of</strong> the pappus in different lots. In our<br />

experiments only one <strong>of</strong> 26 seeds which were released into the air at wind speeds<br />

averaging 7 km per hour and gusting to 22 km per hour became detached from its pappus.<br />

Eighteen out <strong>of</strong> 20 seeds flew out <strong>of</strong> sight at average windspeeds <strong>of</strong> 15 km per hour and<br />

gusting to 38. The pappus <strong>of</strong> one seed became entangled on vegetation and fluttered in<br />

the wind for 15 minutes. but did not separate from the seed (Peschken, unpublished).<br />

Seedlings establish more readily where there is litter. and the moisture in low habitats<br />

may help seedling establishment in pond and ditch margins (Stevens 1926). This is the<br />

reason why S. arvensis is particularly serious in the irrigated area around Outlook,<br />

Saskatchewan (W.J. King. 1980, personal communication). Perennial sow-thistle<br />

thrives on low heavy soils and on lighter soils with a good moisture supply (Stevens<br />

1924).<br />

Fifty-nine shoots per m l reduced the yield <strong>of</strong> oats by 25% (Friesen & Shebeski 1960).<br />

Shaskov el al. (1977) report reduction in wheat yield by 4.5% to 27% in Kazakh SSR. at<br />

3-15 plants per mI. In Saskatchewan five shoots per ml reduced the yield <strong>of</strong> rapeseed by<br />

12% and 10 shoots by 18'Yo (Peschken unpublished).<br />

Control <strong>of</strong> perennial sow-thistle growing in grain is possible with herbicides, but<br />

difficult in broad-leaved crops. Cultivation has to be repeated every three to four weeks<br />

for at least one year to deplete root reserves (Derscheid & Parker 1972).<br />

The annual sow-thistle. S. oleraceus L.. an annual or rarely biennial weed, reproduces<br />

by seeds and has a strong tap root which may reach to a depth <strong>of</strong> 1 m and may<br />

spread to 1.29 m (Kutschera 1960). It grows up to 2 m tall and it produces about 6000<br />

seeds per plant (Salisbury 1961). It is a weed <strong>of</strong> cultivated fields, lawns, gardens. grainfields,<br />

vineyards and occurs in all provinces <strong>of</strong> <strong>Canada</strong>. In <strong>Canada</strong>, it may harbor virus<br />

diseases such as tobacco streak, alfalfa mosaic, cucumber mosaic. and the mycoplasma<br />

disease. aster yellows (Gayed 1978. Holm el al. 1917. Smith 1972). The spiny annual<br />

sow-thistle. S. asper (L.) Hill. is an erect annual or sometimes biennial weed up to 2 m<br />

tall, with a strong taproot reaching down to 2 m (Kutschera 1960). Salisbury (1961)<br />

estimates seed production to be about 18 000 per plant.<br />

205


206 D. P. Pes(<br />

Background<br />

Tephrltis dHscerata<br />

Lw. (Diptera:<br />

Tepbritidae)<br />

Cystiphora sonchl<br />

(Bremi) (Diptera:<br />

Cecidomyiidae)<br />

Both annual sow-thistles occur throughout <strong>Canada</strong>, but are more abundant in Ontario,<br />

Quebec, and British Columbia than in other provinces (Frankton & Mulligan 1970). S.<br />

lenerrimus is only reported from California (Shetler & Skog 1978).<br />

No biological control <strong>of</strong> sow-thistles has been reported from other parts <strong>of</strong> the world. In<br />

<strong>Canada</strong>, only one as yet unidentified Lepidoptera larva has been found feeding on the<br />

flower heads or destroying seeds, and no monophagous insects have been reported<br />

feeding on any other part <strong>of</strong> the perennial sow-thistle (M. Maw, 1982, personal communication).<br />

Conners (1967) lists three disease organisms specialized on S. arvensis:<br />

Marssonina sonchi Dearn. & Bisby, Seploria sonchi-arvensis Deam. & Bisby, and S.<br />

sonchifolia Cke. However, these do not control their host. The survey in eastern Austria<br />

and the Swiss Jura, and <strong>of</strong> the literature by Schroder (1974), and in Iran, Pakistan, and<br />

Japan by Zwolfer (1973) on Sonchus spp. produced only six insect species which appeared<br />

to be sufficiently host specific to warrant further study. Five <strong>of</strong> these are<br />

endophytic in the flower heads and one is a leaf gall fly. Two <strong>of</strong> these, Tephriris dilDcerata<br />

Loew and Cysliphora sonchi (Bremi) have been studied in detail and will be discussed<br />

below. In addition, the moth Celypha rosaceana (Schlager) is reported from Sonchus<br />

spp. and Taraxacum <strong>of</strong>ficinale Weber only and should be studied further (Bradley el al.<br />

1979).<br />

Ecology<br />

The fly deposits clutches <strong>of</strong> six to seven eggs into flower buds and in the laboratory the<br />

oviposition period may last up to ten weeks (Berube 1978a,1978b). The larvae transform<br />

the flower bud into a gall in which they pupate when mature. The adult fly enters<br />

diapause when exposed to a short day or cool temperatures or both and it overwinters as<br />

an adult. There is one generation per year (Berube 1978a). T. dilacerata is widely distributed<br />

in the palaearctic covering about three quarters <strong>of</strong> the range <strong>of</strong> S. arvensis (Berube<br />

1978a).<br />

Ecology<br />

C. sonchi is recorded from Sweden (Sylven 1975) and Denmark in the north (Henriksen<br />

1944) to Italy and Romania (Skuhrav6 el al. 1972) in the south, and from France (Kieffer<br />

1899) in the west to the European part <strong>of</strong> the USSR in the east (Peschken, in press).<br />

C. sonchi females lay their eggs on the underside <strong>of</strong> leaves. Galls are produced and<br />

become about 5 mm in diameter when mature. The larvae pupate in the gall or in the<br />

ground. The adults live up to 16 hours in the laboratory. Up to 721 galls were produced<br />

on one perennial sow-thistle rosette by six females. There are three generations per year<br />

in the field (Skuhrav6 & Skuhravy 1973).<br />

The fly is host specific to Sone/lus spp. (Peschken in press).


Releases and Recoveries<br />

Table 51<br />

Scmc/lus an'ens;s L.. 207<br />

T. dilacerata has been released on a total <strong>of</strong>five sites in Saskatchewan in 1979. 1980. and<br />

1981. and one site each in Alberta. Quebec. and Prince Edward Island (Table 51).<br />

Breeding occurred in 1979 when the flies were released in large field cages (180 cm high.<br />

180 x 180 cm). At Estevan three galls were found in August and 250 at Regina. From the<br />

latter an estimated 1150 flies emerged in the cage at Regina. Many <strong>of</strong> these were allowed<br />

to escape so that they could search for overwintering sites but no survivors were found<br />

in 1980. In 1981. only three galls were found in Outlook and 22 at Wishart. Saskatchewen.<br />

However. 98% <strong>of</strong> the flies were released for overwintering in the fall <strong>of</strong> 1980 and 97% in<br />

1981. All fall releases were made with flies that emerged from galls collected in Austria.<br />

In 1980. the flies emerged in the quarantine laboratory in a 16 hour day except for 1000.<br />

which had emerged in an eight hour day at 14°C and were released into a large field cage<br />

at Regina. Saskatchewan in September 1980. A log pile covered loosely with straw was<br />

provided as a possible overwintering site. The cage was removed in late fall to allow<br />

additional insulation by snow and replaced on 15 April 1981. No flies emerged.<br />

Subsequently it was learned that flies emerging at room temperature in a 16 hour day<br />

do not enter diapause. Releases made in the fall <strong>of</strong> 1981 were all made with flies conditioned<br />

for overwintering.<br />

C. sonciJi was released on four sites in Saskatchewan in 1981. where it survived the<br />

first winter on two sites and produced two generations in 1982 (Table 51). The colony in<br />

Alberta produced galls in 1981 but did not survive the winter. No galls were found from<br />

the 1982 release in Alberta nor from the 1981 and 1982 releases in Manitoba and Quebec.<br />

Open releases and recoveries against Sonchus arvensis L.<br />

Species and Province<br />

Tephritis dilacerala Lw. *<br />

Saskatchewan<br />

Alberta<br />

Quebec<br />

Prince Edward Island<br />

Cysliphora sonchi (Bremi)*<br />

Saskatchewan<br />

Alberta<br />

Manitoba<br />

Quebec<br />

Total<br />

Total<br />

Year Number<br />

1979 810<br />

1980 11486<br />

1981 1093<br />

1981 2000<br />

1981 1947<br />

1981 2000<br />

19336<br />

1981 21900<br />

1981 5000<br />

1982 4500<br />

1982 2000<br />

1981 5000<br />

1982 2100<br />

40500<br />

Year <strong>of</strong> Recovery<br />

1979<br />

1982<br />

1981<br />

* All insects were originally collected in Austria. All C. sonchi and 6% <strong>of</strong> the T. dilacerala<br />

releases were made with laboratory reared stock. and the remaining T. dilacerala releases<br />

were made with flies that emerged from pupae collected in the field in Austria.


208 D. P. Pesch ken<br />

Evaluation <strong>of</strong> Control Attempts<br />

Recommendations<br />

Literature Cited<br />

So far no T. dilaeerata overwintered in Saskatchewan. Saskatchewan, with its thin or<br />

<strong>of</strong>ten non-existent snow-cover, constitutes a harsh climate for overwintering. Also<br />

suitable food for the adult flies may be scarce in spring and fall on the prairies.<br />

(1) Releases <strong>of</strong> CystipllOra sonehi and T. dilaeerata should be made and monitored in<br />

areas where S. arvensis is frequent, i.e. in Ontario and Manitoba, and the northerly areas<br />

<strong>of</strong> Ontario, Quebec. and the Prairie Provinces, and in irrigated areas where S. arvensis is a<br />

problem.<br />

(2) A study <strong>of</strong> the overwintering habits <strong>of</strong> T. dilaeerata should be made.<br />

(3) The role <strong>of</strong> seeds <strong>of</strong> S. arvensis in the propagation <strong>of</strong> existing stands and the<br />

establishment <strong>of</strong> new stands should be investigated to determine whether the establishment<br />

<strong>of</strong> seed-destroying biological control agents should be pursued.<br />

(4) The moth Celypha rosaeeana (Schlager) (Lepidoptera: Tortricidae) and the seedhead<br />

fly Contarinia sellleclllendaliana Rubs. (= soneh; Kieffer) appear to have a narrow<br />

host range and warrant screening as biological control agents.<br />

(5) The survey for S. arvensis insects should be resumed and concentrated in the<br />

northern agricultural areas <strong>of</strong> Europe, such as Sweden, Finland. and the northern areas<br />

<strong>of</strong> European and Asian USSR, where S. arvensis is frequent.<br />

Berube. D.E. (1978a) Larval descriptions and biology <strong>of</strong> Tephritis dilacerata (Dipt.: Tephritidae) a candidate for biocontrol <strong>of</strong> Sonchus<br />

arvensis in <strong>Canada</strong>. Entomophaga 23. 69-82.<br />

Berube. D.E. (1978b) The basis <strong>of</strong> host plant specificity in Tephritis dilacerata and T. formosa (Dipt.: Tephritidae). Entomophaga 23.<br />

331-337.<br />

Boulos. L. (1961) Cytotaxonomic studies in the genus Sonchus. 3. On the cytotaxonomy and distribution <strong>of</strong> Sonchus arvensis L. Botaniska<br />

Notiser 114(1). 57-64.<br />

Bradley. J. D.; Tremewan. W.G.; Smith. A. (1979) British tortricoid moths. Tortricidae: Olethreutinae. Vol. II. London; The Ray <strong>Society</strong> and<br />

British Museum. 320 pp.<br />

Conners. I.L. (1967) An annotated index <strong>of</strong> plant diseases in <strong>Canada</strong>. and fungi recorded on plants in Alaska. <strong>Canada</strong> and Greenland.<br />

Canadian Department <strong>of</strong> Agriculture Publication 1251.381 pp.<br />

Derscheid. L.A.; Parker. R. (1972) Thistles. <strong>Canada</strong> thistle. perennial sow-thistle. South Dakota State University Extension Fact Sheet 450.<br />

4 pp.<br />

Frankton. D.; Mulligan. G.A. (1970) Weeds <strong>of</strong> <strong>Canada</strong>. Canadian Department <strong>of</strong> Agriculture Publication 948. 217 pp.<br />

Friesen. G.; Shebeski. L.H. (1960) Economic losses caused by weed competition in Manitoba grain fields. I. Weed species. their relative<br />

abundance and their eUeet on crop yields. Canadian Journal <strong>of</strong> Plant Science 40. 457-467.<br />

Gayed. S.K. (1978) Tobacco diseases. Canadian Department <strong>of</strong> Agriculture Publication 1641, 57 pp.<br />

Henriksen. K.L. (1944) Fortegne1se over de danske galler (zoodeccidier). Spolia Zoologica Musei Hauniensis VI. 1-212.<br />

Holm. L.G.; Plucknett. D.L.; Pancho. J. V.; Herbcrger.J.P.(1977) Thc world's worst weeds. Honolulu; University Press <strong>of</strong> Hawaii. 609 pp.<br />

Kieffer. J.J. (1899) Enumeration des eecidies recueillies aux Petites-Dalles (Seine-Inferieure) avec description de deux Cccidomyies<br />

nouvelles. Bulletin de la Sociltl des Amis des Sciences Naturelles de Rouen. 89-105.<br />

Kutschera. L. (1960) Wurzelatlas mitteleuropiiischer Ackerunkriiuter und Kulturpllanzen. Frankfurt; DLG. 574 pp.<br />

Pesehken. D.P. (in press) Host specificity and biology <strong>of</strong> Cystiphora sonchi (Loew) (Diptera: Cecidomyiidac). a candidate for the<br />

biological control <strong>of</strong> Sonchus species. Enromophaga.<br />

Salisbury. E. (1976) Weeds and aliens. London; Collins. 384 pp.<br />

SchrOder. D. (1974) The phytophagous insects attacking Sonchus spp. (Compositae) in Europe. Proceedings <strong>of</strong> the 3rd International<br />

Symposium on Biological Control <strong>of</strong> Weeds. Montpellier. France. Commonwealth Institute <strong>of</strong> Biological<br />

Control Miscellaneous Publication 8. 89-96.<br />

Shaskov. V.P.; Kokmakov. P.P.; Volkov. E.E.; Trifanova. L.F. (1977) The influence <strong>of</strong> rhizomatous weeds in spring wheat crops on the<br />

utilization <strong>of</strong> nitrogen. phosphoros and potassium. Agrokhimiya 14(3).57-59.<br />

Shetler. S.G.; Skog. L.E. (1978) A provisional checklist <strong>of</strong> species for Rora North America (Revised). Rora North America Report 84.<br />

Missouri Botanical Garden. 199 pp.<br />

Skuhrava • M.; Skuhravj. V. (1973) Gallmiicken and ihre Gallen auf Wildpllanzen. Wittenberg Lutherstadt; Ziemsen. 118 pp.


Sone/Ills arvensis L., 209<br />

Skuhravd , M.; Skuhravy, v.: Neacsu, P. (1972) Verbreitung der Gallmiicken in Rumlinien. Deutsche entomologische Zeitschrift 19,<br />

375-392.<br />

Smith, K.M. (1972) A textbook or plant virus diseases. 3rd ed., New York; Academic Press, 684 pp.<br />

Stevens, O.A. (1924) Perennial sow-thistle, growth and reproduction. North Dakota Agricultural Experiment Station Bulletin 181,44 pp.<br />

Stevens, O.A. (1926) The sow thistle. NOTlII Dakota Experiment Station Circular 32, 16 pp.<br />

Stevens, O.A. (1932) The number and weight or seeds produced by weeds. Ameriam loumal <strong>of</strong> Botany 19. 784-794.<br />

Sylven, E. (1975) Study on relationships between habits and external structures in Oligotrophidi larvae. (Diptera. Cecidomyiidae).<br />

ZoologiCD Scripta 4, 55-92.<br />

Zwiilrer, H. (1973) A survey ror weed insects in Japan. Iran and Pakistan. Weed projects for <strong>Canada</strong>. Commonwealth Institute <strong>of</strong><br />

Biological Control Progress Report 3D, 27 pp.


Blank Page<br />

210


Pest Status<br />

Biological Control Studies<br />

Discussion<br />

Chapter 44<br />

Verbascum thapsus L., Common Mullein<br />

(Scrophulariaceae)<br />

M.O.MAW<br />

Verbascum thapsus L. is a biennial or, rarely, an annual <strong>of</strong> Eurasian origin. It was<br />

probably introduced into North America several times as a medicinal herb (Gross &<br />

Werner 1978). It is found over most <strong>of</strong> the United States and southern <strong>Canada</strong>, but is rare<br />

on the Canadian prairies, North Dakota, and Montana (Gross & Werner 1978. Maw<br />

1980). It is most vigorous on well drained gravelly or stony soils in rough pastures,<br />

fence rows, roadsides, railway yards, and waste places. It is not a major weed, not<br />

allergenic, not poisonous to humans, and not a problem on cultivated land. It does,<br />

however, harbor insects that may be vectors <strong>of</strong> diseases <strong>of</strong> economic plants. For example,<br />

the mullein leaf bug, Campy/omma verbasci Meyer, whose normal plant food is mullein,<br />

may cause damage to apple or pear fruit and potato leaves (Pickett 1939) and may also be<br />

a vector <strong>of</strong> fire blight, Bacillus amy/ovorus (Burr.) Trev. (Leonard 1965).<br />

The most effective control agents <strong>of</strong> mullein in Europe appear to be a weevil, Gymnaetron<br />

letrum F .• and the mullein shark moth, CucuJ/ia verbasci L. (Noctuidae) (Miotk 1973).<br />

The weevil is already well established in Ontario, British Columbia. and the United<br />

States from New England to Arkansas and the Pacific states (Blatchley & Leng 1916,<br />

Hatch 1971). The adults eat mullein leaves, and the larvae and adults destroy up to 50%<br />

<strong>of</strong> the seeds (Gross & Werner 1978).<br />

C. verbasci was screened (Maw 1980) and although there was some feeding on all<br />

Scrophulariaceae tested, and there was some nibbling on tomato, Nicoliana sp. and<br />

Brass;ca napus, sustained feeding and development occurred only on Verbascum spp.<br />

V. thapsus is an early colonizer on abandoned agricultural fields with the tall flowering<br />

stocks usually appearing the second year after abandonment. Populations are usually<br />

scattered but locally abundant, and the weed may cover entire fields where the ground<br />

has been left undisturbed. This situation is, however, relatively short-lived. In a study<br />

by Gross & Werner (1978) it was found that in a barley field that had been abandoned for<br />

three years, the density <strong>of</strong> V. thapsus was 1.0 plants per mI. In an adjacent field that had<br />

been abandoned for 12 years, the density <strong>of</strong> V. thapsus had declined to O.17/m l • Since<br />

seeds can be dormant but viable for 35 years, the typical pattern seems to be one <strong>of</strong><br />

ephemeral adult populations and long-lived seed pools (Gross & Werner 1978).<br />

Although the mullein leaf bug. Campy/omma verbasci, may damage apple and pear<br />

fruit, the bug is only partially phytophagous. It cannot complete its life cycle without<br />

insect prey and feeds on several species <strong>of</strong> mites and insects that attack the fruit crop<br />

(Gross & Werner 1978).<br />

211


212 M. G. Maw<br />

Recommendations<br />

Literature Cited<br />

The mullein moth, Cucullia verbasci, is considered to be a safe biological control<br />

agent to release. However, the inherent risk in releasing any organism into newenvironmental<br />

conditions remains and until the economic value in controlling the weed is<br />

determined, no release should be made.<br />

Until the loss caused by mullein has been determined, no releases <strong>of</strong> biological control<br />

agents should be made.<br />

Blatchley, W.S.; Leng, C.W. (1916) Rbynchophora or weevils <strong>of</strong> North Eastern America. Indianapolis; Nature Publishing Co., 682 pp.<br />

Gross, K.L.; Werner, P.A. (1978) The biology <strong>of</strong> Canadian weeds. 28. Verboscum thopsus L. and V. blal/aria L. Canadian Journal <strong>of</strong><br />

Plant Science 58, 401-413.<br />

Hatch, M.H. (1971) The Beetles <strong>of</strong> the Pacific Northwest. Part V. University <strong>of</strong> Washington Press, 662 pp.<br />

Leonard, D.E. (1965) Atractotomus mali and Campylomma verbosci (Heteroptera: Miridae) on apples in Connecticut. Journal <strong>of</strong><br />

Economic Entomology 58(5),1031.<br />

Maw, M.G. (1980) Cucullia verbose;, lin agent for the biological control <strong>of</strong> common mullein (Verboscum thapsus). Weed Science 28,<br />

27-30.<br />

Miotic, P. (1973) Phytophagous insects associated with weeds in central Europe. Part II. Weed Projects for <strong>Canada</strong>. Commonwealth<br />

Institute <strong>of</strong> Biological Control Progress Report 32, 29 pp.<br />

Pickett, A.D. (1939) The mullein (eafbug, Campylomma verbose; Meyer, as a pest <strong>of</strong> apples in Nova Scotia. <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> Ontario<br />

Annual Report 69, 105-106.


PART III<br />

BIOLOGICAL CONTROL OF FOREST<br />

INSECT PESTS IN CANADA 1969-80


Blank Page<br />

214


216 M. A. Hulme and G. W. Green<br />

Selection <strong>of</strong> Biological Control Agents<br />

pusilla (Lep.), both <strong>of</strong> which feed on birch; the gypsy moth, Lymamria dispar (L.); the<br />

balsam fir sawfly, Neodiprion abietis (Harr.), which, in the first review, was only examined<br />

using surplus parasitoid stock; Bruce spanworm, Operophtera bruceata (Hulst); two<br />

tussock moths, Orgyia leucostigma (J .E. Smith) and Orgyia pseudotsugata (McDunnough);<br />

and the mountain-ash sawfly, Pristiphora geniculata (Htg.). Effort was almost equally<br />

divided between treatments with entomopathogens and releases <strong>of</strong> predators or parasitoids.<br />

The current review thus contains evaluations <strong>of</strong> control attempts against 21 insect<br />

pests, although in some cases this simply entails updating assessments <strong>of</strong> control<br />

attempts made during the previous review period.<br />

Despite the increased number <strong>of</strong> pests covered in the current review, total effort<br />

devoted to biological control <strong>of</strong> forest pests declined as the Canadian Forestry Service<br />

went through a period <strong>of</strong> austerity. Work with predators and parasitoids suffered<br />

particularly severe reductions <strong>of</strong> resources; and certain aspects <strong>of</strong> all programmes, such<br />

as the population dynamics <strong>of</strong> the pest and <strong>of</strong> the control agent, were almost totally<br />

neglected. These cutbacks are reflected in the lack <strong>of</strong> depth with which many <strong>of</strong> our<br />

investigations were undertaken and inevitably reduced our chances <strong>of</strong> finding successful<br />

controls. Despite these difficulties significant progress has been made, as detailed<br />

later.<br />

The definition <strong>of</strong> biological control <strong>of</strong> forest insect pests used in this review follows<br />

that <strong>of</strong> the previous review - it is confined to the regulation <strong>of</strong> pest populations by the<br />

introduction <strong>of</strong> parasitoids, predators, or entomopathogens. Population regulation is<br />

considered acceptable if damage is controlled at tolerable levels in economic andlor<br />

·0 social terms. The methods <strong>of</strong> introduction <strong>of</strong> the control agent cover all strategies from<br />

inoculative releases <strong>of</strong> an organism, which then spreads through the pest population, to<br />

augmenting agents already present in the pest popUlation, or to inundation <strong>of</strong> the pest's<br />

environment with a control agent that mayor may not already be present (Knipling<br />

1979).<br />

The chapters that follow this overview are organized in alphabetical generic sequence,<br />

and each chapter is written by the scientists responsible for conducting or assessing the<br />

control attempts. An outline is given <strong>of</strong> the present status <strong>of</strong> the pest, background is<br />

given where needed on reasons for the choice <strong>of</strong> control agents, and the authors'<br />

evaluation is given <strong>of</strong> each attempt. Each chapter concludes with specific suggestions<br />

for future work.<br />

In describing the pest status, and in many <strong>of</strong> the assessments <strong>of</strong> control effectiveness<br />

reference is made to the Forest Insect and Disease Survey (FIDS), an organization<br />

within the Canadian Forestry Service that carries out major surveys <strong>of</strong> pest populations<br />

and pest damage across the country. In earlier times surveys included routine rearings to<br />

measure parasitoid levels in pest populations, but due to staff cutbacks such measurements<br />

are rarely performed now and surveys by FIDS are generally confined to measurements<br />

<strong>of</strong> pest numbers and pest damage. This in turn has seriously impeded the<br />

effectiveness <strong>of</strong> the biological control research applied to forest pests.<br />

For taxonomic and other reasons, many changes in the names <strong>of</strong> insects or the spelling<br />

<strong>of</strong> insect names have taken place during the current review period. In order to standardize<br />

sQme <strong>of</strong> these changes, forestry contributions follow the nomenclat.ure <strong>of</strong> insect pests<br />

given in the "Common names <strong>of</strong> insects and related organisms" published by the<br />

<strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> America in 1980.<br />

Inoculative releases <strong>of</strong> parasitoids and predators<br />

When planning this strategy <strong>of</strong> biological control it has become customary to first<br />

establish whether the target pest is a native or an introduced species. Of the 21 pests


Biological control <strong>of</strong> forest insect pests 217<br />

included in this review, 9 are considered native: C. fum;ferana, spruce budworm; C.<br />

occidentalis, western spruce budworm; M. disslria, forest tent caterpillar; N. abielis.<br />

balsam fir sawfly; N. lecontei, redheaded pine sawfly; N. swainei, Swaine jack pine<br />

sawfly; Operophtera bruceala. Bruce spanworm; Orgyia lellcostigma. whitemarked<br />

tussock moth; and O. pselldot.mgata. Douglas-fir tussock moth. There is disagreement on<br />

whether P. erichsonii. larch sawfly, is native or introduced. although most favour the<br />

latter view (A. Giroux' 1981. personal communication; Pschorn-Walcher 1977). Following<br />

this division into introduced and native species it has been customary to concentrate<br />

on introduced pests as being promising candidates for biological control by inoculative<br />

introduction <strong>of</strong> predators or parasitoids obtained from the pest's native habitat. Thus. in<br />

the present review. all but two such control attempts are against introduced pests: the<br />

exceptions are attempts against C. fumiferatla. where mostly limited and unpromising<br />

cage studies were formed with parasitoids. and attempts against N. swainei. where trials<br />

were made to introduce Formica spp. <strong>of</strong> ants as predators. The former attempts are best<br />

described as preliminary investigations rather than control attempts. and the latter trials<br />

as introductions <strong>of</strong> polyphagous predators aimed at control <strong>of</strong> several forest pests rather<br />

than just the sawfly. In general. the number <strong>of</strong> parasitoids or predators released in each<br />

control attempt depended on numbers obtained from field collections. Because <strong>of</strong><br />

limited resources <strong>of</strong> both people and material. little rearing was carried out to increase<br />

insect numbers and few <strong>of</strong> the female parasitoids were deliberately mated before<br />

release. Reeks & Cameron (1971) feel that a minimum <strong>of</strong> 500 mated females should be<br />

released but this was rarely achieved in the work reported here. Indeed. Beirne (1975)<br />

regards releasing inadequate numbers and other faulty procedures as important reasons<br />

for failure in control attempts.<br />

The question <strong>of</strong> whether to release one or several species <strong>of</strong> parasitoids or predators<br />

continues to evoke controversy (e.g. DeBach el al. 1976). Both release strategies were<br />

practised during the current review period without either showing obvious superiority in<br />

terms <strong>of</strong> control success. In co-operation with CIBC Delemont and other shipping<br />

stations, precautions were taken to screen out unwanted parasitoids, such as c1eptoparasitoids<br />

and hyperparasitoids. before releases were attempted. A further difficult question<br />

is whether to release monophagous or polyphagous parasitoids. In the case-studies<br />

reported here both types shared in the control successes. In all, 31 species <strong>of</strong> parasitoids<br />

and predators were released against 11 target pests and a further 3 species <strong>of</strong> parasitoids<br />

were tested in laboratory and field cage studies against C. fumiferana. Well over half <strong>of</strong><br />

the introductions were from Europe, three were from Japan. and one from Argentina;<br />

the remainder were relocations within <strong>Canada</strong>. All but three <strong>of</strong> the species belonged to<br />

the order <strong>of</strong> Hymenoptera; the three exceptions. all Diptera. were a laboratory culture <strong>of</strong><br />

Agria house; Shewell tested unsuccessfully against C. fumiferana, Lypha dubia<br />

Fallen, released against R. buoliana but not known to be established, and<br />

Cyzenis albicans (Fall.) successfully established and helping to control O. brumala.<br />

Approximately half <strong>of</strong> the released parasitoids and predators are known to be established<br />

and the list may be extended as more time is allowed for establishment to become evident.<br />

A summary <strong>of</strong> all free field releases is shown in Table 52. Three parasitoids tested only in<br />

cages are also included to complete the list <strong>of</strong> those examined during the review period.<br />

Application <strong>of</strong> entomopathogens<br />

Entomopathogens were used for attempted control against all the native insect pests<br />

covered in this review. and similar controls were explored against the introduced pests<br />

N. serlifer and L. dispar. The main entomopathogens used here were the bacterium B.I.<br />

and a number <strong>of</strong> baculoviruses.<br />

, Biosystematics Research Institute, Ottawa


218 M. A. Hulme and G. W. Green<br />

Table 52 Free field releases <strong>of</strong> parasitoids and predators against forest insect pests between 1969<br />

and 1980<br />

Host<br />

Chorisloneura fumiferana<br />

Coleophora laricella<br />

Coleophora serralella<br />

Fenusa pusilla<br />

Lymantria dispar<br />

Neodiprion seTlifer<br />

Neodiprion swainei<br />

Operophlera brumala<br />

Prisliphora erichsonii<br />

Prisliphora geniculala<br />

Rhyacionia buoliana<br />

• Laboratory culture .<br />

•• Cage release only.<br />

Parasitoidslpredators<br />

Agria housei<br />

Cephaloglypla laricis<br />

Cephaloglypla murinanae<br />

Lissonola sp.<br />

Trichogramma sp.<br />

Agalhis pumila<br />

Chrysocharis laricinellae<br />

Diadegma laricinellum<br />

Dicladocerus japonicus<br />

Apanteles coleophorae<br />

Apanleles mesoxanthus<br />

Apanteles corvin us<br />

Campoplex borealis<br />

Campoplex sp.<br />

GrypocentTUS albipes<br />

Lalhrolesles nigricollis<br />

Anaslaws dis paris<br />

Ooencyrlus kuvanae<br />

Dipriocampe diprioni<br />

Exenterus abruplorius<br />

Lophyropleclus IUlealor<br />

Pleolophus basizonus<br />

Formica lugubris<br />

Formica obscuripes<br />

Agrypon flaveolawm<br />

Cyzenis albicans<br />

Mesoleius lenthredinis<br />

Olesicampe benefaclor<br />

Olesicampe geniculalae<br />

Rhorus sp.<br />

Lypha dubia<br />

Orgilus obscuralor<br />

Parasierola nigrifemus<br />

Agathis binominala<br />

I<br />

Total<br />

number<br />

released<br />

2800 •<br />

••<br />

••<br />

59<br />

20000··<br />

2633<br />

530<br />

101<br />

292<br />

337<br />

1480<br />

1 753<br />

4994<br />

1404<br />

25000<br />

1 189<br />

4072<br />

2389<br />

1032<br />

1000 000<br />

5000000<br />

820<br />

1238<br />

467<br />

8588<br />

912<br />

63<br />

496<br />

560<br />

595<br />

85<br />

B.I. has been tried against a variety <strong>of</strong> pests, all belonging to the Lepidoptera, because<br />

the pathogenic effects <strong>of</strong> many strains <strong>of</strong> the bacterium are promoted by the alkaline<br />

digestive conditions typical <strong>of</strong> this order <strong>of</strong> insects. Much <strong>of</strong> the pioneering work on the<br />

bacterium was carried out by the Canadian Forestry Senice and work on the mode <strong>of</strong><br />

action <strong>of</strong> the entomopathogen is continuing. B.I. is now used operationally by forest<br />

managers: the control agent is fully registered as a pesticide in <strong>Canada</strong> and is commercially<br />

produced in a number <strong>of</strong> different formulations for use against C. fumiferana, L. dispar,<br />

and several other lepidopterous defoliators. Methods <strong>of</strong> applying the bacterium have<br />

varied. In a few cases the bacterium was applied from ground sprays on localized areas


Biological control <strong>of</strong>forest insect pests 219<br />

<strong>of</strong> pest outbreaks (examples are treatments to control M. disstria and L. dispar) ,<br />

whereas in most other cases the bacterium was applied with aerial sprays (examples<br />

here include attempted control <strong>of</strong> C. fumiferana, C. occidentalis, and O.<br />

pseudotsugata). Most development <strong>of</strong> aerial spraying <strong>of</strong> B.t. was conducted with C.<br />

fumiferana. Both water-based and oil-based formulations were used and the spray<br />

aircraft included fixed-wing aircraft equipped with from one to four engines, and<br />

various sizes <strong>of</strong> helicopters. Areas treated ranged as high as 90 000 ha, and in all cases<br />

results to a large extent depended on the application techniques that were employed.<br />

The time period during which B.t. can be applied during the insects' development is<br />

shorter than that permissible with chemical insecticides because the bacterium must be<br />

ingested to be effective whereas the standard chemical insecticides are neurotoxins and<br />

exhibit contact toxicity. An interesting international aspect <strong>of</strong> this work involved cooperative<br />

trials between <strong>Canada</strong> and the United States under the 6-year CAN USA<br />

Spruce Budworms Research Programme that started in 1977. The two countries also cooperated<br />

in one trial using B.t. against O. pseudotsugala in British Columbia.<br />

The use <strong>of</strong> viruses, particularly baculoviruses, for control <strong>of</strong> forest insect pests has<br />

progressed rapidly since the last review, and Morris (1980) provides a convenient<br />

tabulaton <strong>of</strong> all field tests where either ground or aerial applications <strong>of</strong> virus were used.<br />

Most attention has been focused on the neodiprionid sawflies because NPVs that attack<br />

these sawflies were found to be particularly virulent. Many sawflies also feed colonially,<br />

which facilitates cross infection, and some <strong>of</strong> the viruses spread from the point <strong>of</strong><br />

infection, thus allowing spot introductions <strong>of</strong> virus to be used rather than complete<br />

coverage <strong>of</strong> the infected areas. Predacious and scavenging insects are thought to be the<br />

main disseminators <strong>of</strong> the viruses. In some cases strips <strong>of</strong> the infected area were<br />

sprayed, and in the case <strong>of</strong> N. swainei spot introductions <strong>of</strong> laboratory-infected pupae<br />

were used. Other viruses, again mainly NPVs, with good virulence against a number <strong>of</strong><br />

Lepidoptera, have also been evaluated. Examples are tests against the tussock moths,<br />

O. leucostigma, and O. pseudotsugata. Some less virulent viruses were also applied<br />

aerially against, for example, C. fumiferana and C. occidenlalis; others were applied by<br />

ground sprays. for example. those against M. disstria and L. dispar.<br />

The remaining entomopathogens covered in this review are essentially at the laboratory<br />

stage <strong>of</strong> development. Fungi are well known to cause epizootics, such as the one<br />

believed to have helped collapse a recent outbreak <strong>of</strong> hemlock looper, Lambdina<br />

fiscellaria fiscellaria (Guenee), in Newfoundland (Otvos 1973). Other North American<br />

forest pests in which fungal epizootics have been recorded include spruce budworm C.<br />

fumiferana, black headed budworm, Acleris variana (Fern), and forest tent caterpillar,<br />

M. disslria. Entomophthora spp. <strong>of</strong> fungi are generally responsible and these are the<br />

fungi on which the Canadian Forestry Service concentrates its effort. Much <strong>of</strong> the<br />

biology <strong>of</strong> these fungi is now elucidated; before direct applications are attempted, efforts<br />

continue to devise techniques for mass production <strong>of</strong> resting spores, to devise practical<br />

application systems, and to induce spore germination, particularly under field conditions.<br />

One study, not reported elsewhere in this review, is an examination <strong>of</strong> control possibilities<br />

<strong>of</strong> mountain pine beetle, Dendroctonus ponderosae Hopkins, with Beauveria bassiana<br />

Vuillemin. This work can be summarized by stating that although the fungus is lethal to<br />

the insect in laboratory trials, no satisfactory way has yet been found to infect insects in<br />

the field (S. Whitney2 1981, personal communication).<br />

Work with protozoa roughly parallels that with fungi and emphasis is now on epizooticlogical<br />

studies. Protozoa are dominant entomopathogens <strong>of</strong> insects such as spruce<br />

budworm, C. fumiferana, but their effects are to debilitate rather than to kill the insects,<br />

by affecting larval and pupal vigour and by reducing adult longevity and fecundity.<br />

Because these entomopathogens are ubiquitous. work continues to elucidate their<br />

interaction with other control agents.<br />

2 Pacific Forest Research Centre, Victoria. British Columbia.


220 M. A. Hulme and G. W. Green<br />

Evaluation <strong>of</strong> Control Attempts<br />

Table 53<br />

Virtually no work has been carried out with nematodes during the review period,<br />

although recent laboratory studies have shown that Heterorhabditis hefiothidis (Khan et<br />

al.) is lethal to C. fumiferana (G. Finney' 1981 personal communication). Work with<br />

rickettsia has been deliberately avoided because <strong>of</strong> potential human health problems.<br />

We have attempted to summarize the authors' evaluations <strong>of</strong> control success under the<br />

headings <strong>of</strong> predators and parasitoids, and <strong>of</strong> entomopathogens. As previous reviews in<br />

this series have been devoted largely to predators and parasitoids, we have also<br />

summarized these earlier evaluations <strong>of</strong> the major control attempts to show progress<br />

through the review periods. Simmonds' classification <strong>of</strong> control success has been used<br />

to provide continuity with the previous review (Simmonds 1969). Practical control using<br />

entomopathogens received relatively little attention until the current review period,<br />

except for work with NPVs against G. hercyniae and N. lecontei, hence few comparisons<br />

can be made with earlier work. Furthermore, Simmonds' classification <strong>of</strong> control<br />

success does not adequately describe the control strategy attempted with many entomopathogens<br />

and a separate rating system has thus been devised for entomopathogens that<br />

better reflects the behaviour <strong>of</strong> these control agents.<br />

The summary <strong>of</strong> control attempts given in Tables 53 and 54 shows that two-thirds <strong>of</strong><br />

the forest insect pests covered in this review were considered to be successfully<br />

Evaluation <strong>of</strong> biological control attempts against forest insect pests using predators and<br />

parasitoids<br />

Degree <strong>of</strong> success·<br />

Pest 1910-58 1958-68 1969-80<br />

Adelges piceae promising<br />

Coleophora laricella good ++++ ++++<br />

Coleophora se"atella<br />

Fenusa pusilla +<br />

Gilpinia hercyniae good ++++*. ++++**<br />

Leucoma salicis good +++ +++<br />

Lymantria dispar<br />

Neodiprion sertifer ++ ++<br />

Operophtera brumata ++++ ++++<br />

Pristiphora erichsonii promising ++ +++<br />

Pristiphora geniculata ++<br />

Rhyacionia buoliana + +++<br />

• Based on McGugan & Coppel's assessment up to 1958, and on Reeks & Cameron's<br />

evaluation for 1959-68. Simmonds' (1969) classification <strong>of</strong> control success is used<br />

in the final two columns, viz.<br />

No control.<br />

+ Slight pest reduction or too early for evaluation <strong>of</strong> control.<br />

+ + Local control; distribution restricted or not fully investigated.<br />

+ + + Control widespread but local damage occurs.<br />

+ + + + Control complete<br />

** Acting in conjunction with a virus.<br />

) Memorial University <strong>of</strong> Newfoundland, St. Johns, Newfoundland.


Table 54<br />

Biological control <strong>of</strong> forest insect pests 221<br />

Evaluation <strong>of</strong> biological control attempts against forest insect pests using entomopathogens·<br />

Entomopathogens<br />

Pest Type Degree <strong>of</strong> success··<br />

Chorisloneura fumiferana NPV-CPV-GV-EV·" slight<br />

Chorisloneura occidenraiis<br />

GiJpinia herr:yniae<br />

Lyl11lJlJJria dispar<br />

Maiacosoma disstria<br />

Neodiprion abietis<br />

Neodiprion Iecontei<br />

Neodiprion sertifer<br />

Neodiprion swainei<br />

Operophtera bruceala<br />

Orgyitl Jeucostigma<br />

Orgyitl pseudotsugata<br />

••<br />

...<br />

••••<br />

8.1.<br />

NPV<br />

8.1.<br />

NPV<br />

NPV<br />

8.1.<br />

NPV<br />

8.1.<br />

NPV<br />

NPV<br />

NPV<br />

NPV<br />

NPV<br />

NPV<br />

NPV<br />

B.t.<br />

good<br />

none<br />

none<br />

exceUent· .. •<br />

slight<br />

slight<br />

none<br />

good<br />

unknown<br />

excellent<br />

excellent<br />

good<br />

unknown<br />

good<br />

good<br />

slight<br />

Covering the period 1969-80, except for N. Ieconrei and G. herc:yniae where work<br />

was undertaken before 1969.<br />

The rating system is as follows:<br />

slight - decreased pest population but little damage prevented.<br />

good - decreased pest population and damage reduced to economically acceptable<br />

levels<br />

excellent - pest populations and damage reduced to almost zero.<br />

NPV<br />

CPV<br />

GV<br />

EV<br />

B.t.<br />

nuclear polyhedrosis virus<br />

cytoplasmic polyhedrosis virus<br />

granulosis virus<br />

entomopox virus<br />

Bacillus thuringiensis<br />

Acting in conjunction with parasitoids .<br />

regulated by using biological control agents and that success was equally divided<br />

between parasitoids and entomopathogens. A number <strong>of</strong> evaluations cannot yet be<br />

made due to lack <strong>of</strong> data, although in some cases, as mentioned later, prospects for<br />

success are promising. Brief details <strong>of</strong> control evaluations with parasitoids and predators,<br />

and with entomopathogens are as follows.<br />

Parasitoids and predators<br />

Many <strong>of</strong> the following chapters report work continued from the previous review period.<br />

Most evaluations have been directed to technical success in control and little further can<br />

be added to the earlier tentative economic assessments <strong>of</strong> Reeks & Cameron (1971). C.


222 M. A. Hulme and G. W. Green<br />

laricella, G. hercyniae. L. salicis. O. brumata, and P. erichsonii continue to be controlled<br />

by introduced parasitoids (Table 53), as does R. buoliana, provided temperatures favour<br />

parasitoids by synchronising their development with nectar and pollen supplies. These<br />

supplies are provided by encouraging the growth <strong>of</strong> flowering plants such as buckwheat<br />

(Fagopyrum esculentum Moench), wild carrot (Daucus carota L.). and milkweed (Asclepias<br />

syr;aca L.). This effective and novel approach emphasizes the importance <strong>of</strong> examining<br />

several alternative tactics for inoculating and establishing the control agent in the pest<br />

population. The geographic range <strong>of</strong> control <strong>of</strong> some <strong>of</strong> the above insect pests does not yet<br />

cover the entire country and attempts are now being made to control, for example. C.<br />

laricella in western <strong>Canada</strong> by introducing a braconid, an ichneumonid and two eulophids<br />

similar to those that are successful in eastern <strong>Canada</strong>.<br />

Several new subjects were chosen for control attempts. Small introductions <strong>of</strong> braconids<br />

and ichneumon ids were made to control C. serratella, so far without success.<br />

However, prospects still seem good because these parasitoids are important regulators<br />

<strong>of</strong> populations in Europe. Limited introductions were made against F. pusilla and at<br />

least one <strong>of</strong> the parasitoids has become established, although it is too early to evaluate<br />

control potential. P. geniculata was chosen as a promising candidate for control because<br />

it is so heavily parasitized in its native Europe that once sufficient numbers <strong>of</strong> the<br />

ichneumonid were received from Europe good local control was obtained. underlining<br />

the importance <strong>of</strong> making adequate releases. Attempts against C. fumiferana are not<br />

evaluated here because they essentially did not progress beyond cage studies. Attempts<br />

against N. swaine; are also not evaluated because they involved release <strong>of</strong> a general<br />

predator <strong>of</strong> many forest pests.<br />

Entomopathogens<br />

Entomopathogens have been used for most <strong>of</strong> our current biological control attempts<br />

against native pests, mainly by attempting to inundate the environment <strong>of</strong> the pest<br />

insect. In general. entomopathogens must be applied early in the insect's development if<br />

foliage is to be saved in the year <strong>of</strong> application. Early application is particularly important<br />

when viruses are used.<br />

Remarkable progress has been made with B.t. during the review period. Improved<br />

strains <strong>of</strong> B.t. (Dulmage 1970) followed by improvements in formulation and application<br />

technology have now reached the point where many forest managers find B.I. can be as<br />

effective as chemical insecticides against C. fumiferana (e.g. Dorais el al. 1980); costs<br />

are three to five times higher. It is. <strong>of</strong> course, essential that the material be applied at the<br />

correct time, in adequate dosage, and with good application techniques that include<br />

suitable formulation <strong>of</strong> the spray ingredients. Many <strong>of</strong> these developments with B.I.<br />

sprays have not yet been adequately tested on other Lepidoptera but there are no a<br />

priori reasons why success should not be similar.<br />

Recent success with viruses has been outstanding. Effective viruses were found<br />

against all the tested neodiprionid sawflies, with the possible exception <strong>of</strong> N. ab;etis,<br />

where sufficient data are not yet available to make an assessment. The NPV <strong>of</strong> N.<br />

leconlei has been particularly successful in spreading from the point <strong>of</strong> application<br />

(apparently via predacious and scavenging insects) and has allowed spot introductions<br />

to be used. Observations suggest that many applied viruses continue to exert control in<br />

the years following application, although it is not yet clear in many cases how this carryover<br />

is achieved. The virus may infect <strong>of</strong>fspring transovarially, for example. or the virus<br />

may simply contaminate the environment by residing in cadavers. twig crotches, and<br />

other protected locations. Further research is required to elucidate the specific mechanisms.<br />

Effective viruses were also found that could control certain Lepidoptera, notably the<br />

NPVs for O. leucostigma and O. pseudotsugata. Success against C. fum;ferana and L.<br />

dispar has been only marginal because <strong>of</strong> the lower virulence <strong>of</strong> the viruses tested.


Future Considerations<br />

Biological control <strong>of</strong> forest insect pests 223<br />

The estimated cost for virus material largely depends on production methods. Thus<br />

NPV from field-infected sawflies costs about $2.50Iha, whereas the tussock moth NPV<br />

produced in vivo in the laboratory costs about S501ha.<br />

There have been no control successes with either fungi or protozoa during the review<br />

period because, with the exception <strong>of</strong> trials to infect D. ponderosae with B. bassiana<br />

mentioned earlier, direct control has not been attempted. Basic questions in epizootiology<br />

remain to be answered first. Protozoa are, in any case, not generally envisaged as<br />

satisfactory control agents by themselves and emphasis is directed to control strategies<br />

in which they act synergistically with other agents.<br />

Like earlier reviews, the present reports cover many approaches to biological control.<br />

These case histories provide a valuable base on which to build future work, whether it be<br />

from the standpoint <strong>of</strong> assessing development with specific control agents or <strong>of</strong><br />

examining control opportunities with specific insect pests. Each <strong>of</strong> these aspects will be<br />

considered separately in the following discussion.<br />

Predators and parasitoids as control agents<br />

There are two schools <strong>of</strong> thought on whether to emphasize introduced pests in inoculative<br />

control attempts with exotic parasitoids or predators. Although it is agreed that this<br />

approach has been fruitful for many introduced pests, some favour a continuation <strong>of</strong> this<br />

planning strategy, whereas others such as Munroe (1971) point out that this concentration<br />

<strong>of</strong> effort on introduced pests has no theoretical basis to support it. Indeed, Pimentel<br />

(1963) lists a number <strong>of</strong> successes against native pests. One interesting recent addition to<br />

such lists is Oxydia trychiata Gn., a native forest defoliator in Colombia that was<br />

controlled by Telenomus a/sophilae Vier., an egg parasitoid <strong>of</strong> A/sophila pometaria<br />

(Harris) imported from Virginia in the United States. Not only does this example show<br />

control <strong>of</strong> a native pest by a parasitoid from a host <strong>of</strong> a different genus but it also<br />

demonstrates a successful transfer from temperate to tropical latitudes (Bustillo &<br />

Drooz 1977). The example serves to underline the earlier dictum <strong>of</strong> Huffaker et al. (1971)<br />

that "native as well as exotic pests are suitable subjects for biological control importations".<br />

Indeed we hope that inoculative introductions <strong>of</strong> exotic parasitoids and predators<br />

against native forest pests will be given more attention than in the past.<br />

A further point that perhaps needs to be reconsidered is the method by which<br />

parasitoid species are selected. Disproportionate attention has been devoted to the<br />

Hymenoptera during the review period, yet among other orders the Diptera, in particular,<br />

contain numerous species <strong>of</strong> successful parasitoids that deserve closer examination.<br />

Diptera can be more difficult to rear and the adults are sometimes considered more<br />

difficult to identify than Hymenoptera, but these factors should not prejudice the choice<br />

<strong>of</strong> candidate parasitoids unless all other comparative factors are judged equal. Munroe<br />

(1971) pointed out that during the previous review period the establishment rate for<br />

introduced Diptera was actually higher than that for introduced Hymenoptera although<br />

it is, <strong>of</strong> course, recognized that control does not always follow establishment.<br />

In terms <strong>of</strong> the number <strong>of</strong> species released we concur with the views <strong>of</strong> Reeks &<br />

Cameron (1971) that either single or multiple species' releases are acceptable when<br />

supporting work shows that reasons are sound. Case histories suggest even where<br />

competitive displacement occurs, the resulting control may well be improved (e.g.<br />

Caltagirone 1981). Precautions should, <strong>of</strong> course, be taken to screen out undesirable<br />

parasitoids, especially hyperparasitoids and c1eptoparasitoids (e.g. Schroeder 1974).<br />

Desirable parasitoids should have good numerical andlor functional response to host<br />

population changes through attributes such as high fecundity, efficient host searching


224 M. A. Hulme and G. W. Green<br />

ability, and good synchronization <strong>of</strong> reproduction with the host (Pschorn-Walcher<br />

1977). Inbreeding that restricts genetic variability should be limited. It is always advisable to<br />

try to ensure genetic variation between released individuals by collecting breeding stock<br />

from several locations (Bennett 1974, Mackauer 1976).<br />

Selection <strong>of</strong> collecting locations for exotic parasitoids and predators has rightly been<br />

focused on the native habitat <strong>of</strong> the pest when introduced insects are the subject <strong>of</strong><br />

control attempts and this has inevitably led to a concentration <strong>of</strong> collection efforts in<br />

Europe. Case histories, however. show that successful control agents have also been<br />

found on different hosts and in different parts <strong>of</strong> the world. and to allow for this<br />

possibility. and thus increase the probability <strong>of</strong> success, a proportion <strong>of</strong> effort should be<br />

devoted to collecting from areas other than the native habitat <strong>of</strong> the pest (Pschorn­<br />

Walcher 1977). This approach is, <strong>of</strong> course. essential where native pests are the subject<br />

<strong>of</strong> control attempts by inoculative releases.<br />

Although the above remarks are aimed principally at so-called classical biological<br />

control. they are not intended to preclude attention to other release strategies. Augmentative<br />

or inundative releases <strong>of</strong> parasitoids or predators should perhaps be considered<br />

in some cases. These techniques remain essentially untested in Canadian forestry<br />

applications despite successes claimed in Russia and China (Tropin et 01. 1980, McFadden<br />

et al. 1981). For example, in Russia release <strong>of</strong> Trichogromma evanescens Westw. against<br />

eggs <strong>of</strong> R. buoliana and Petrona resine/la (L.) reduced the number <strong>of</strong> damaged pine buds<br />

by over 50% (Beglyarov & Smetnik 1977); and in China Trichogrommo dendrolimi Mats.<br />

reduced infestation by Dendrolimus sibericus Tschetw. from 63% <strong>of</strong> trees in 1956 to 1 %<br />

<strong>of</strong> trees in 1970 (Hussey 1977). Attention could also be given to the use <strong>of</strong> behavioural<br />

chemicals such as kairomones (e.g. Gross Jr. 1981, Weseloh 1981, Vinson 1977); these<br />

semiochemicals should improve control by augmenting the ability <strong>of</strong> natural enemies to<br />

locate hosts.<br />

Finally, two fundamental research needs in predator and parasitoid introductions<br />

have been raised repeatedly in this overview. The first need covers rearing techniques to<br />

ensure that adequate numbers <strong>of</strong> biological control organisms (mated females) are available<br />

for whatever release strategy is employed. The second fundamental need is for better<br />

knowledge <strong>of</strong> the population dynamics <strong>of</strong> the pest insect and <strong>of</strong> the candidate predators<br />

or parasitoids. Many feel that neglecting this aspect in favour <strong>of</strong> an ad hoc approach<br />

inevitably reduces chances <strong>of</strong> obtaining successful control.It is <strong>of</strong> course recognized. as<br />

Pschorn-Walcher (1977) states, that "predictions derived from multivariate analysis or<br />

population models are no guarantee that the biological control agents selected will be<br />

effective." Judgement must be exercised in deciding how much background information<br />

is desirable before releases are attempted. As Simmonds (1972) points out. "The introduction<br />

... is the crucial experiment. Promising biological control agents have failed to<br />

live up to expectations whereas apparently unlikely species have been very successful".<br />

Entomopathogens as control agents<br />

Bacteria and viruses have dominated control successes with entomopathogens and will<br />

probably continue to do so in the immediate future as improved methods are found for<br />

producing and applying these organisms. With B.t .• for example, efficacy against C.<br />

fumi/erona has been widely demonstrated and future research emphasis will be on<br />

reducing costs. This can be approached in a number <strong>of</strong> ways. Development <strong>of</strong> more<br />

concentrated formulations to reduce the volume <strong>of</strong> carrier liquid will reduce both shipping<br />

and application costs. Development <strong>of</strong> even more potent strains <strong>of</strong> B.t. will lead to similar<br />

reductions in costs. Finally, development <strong>of</strong> better application methods will increase the<br />

probability that larvae will ingest a lethal dose from a given quantity <strong>of</strong> applied B.t. by<br />

optimising the distribution <strong>of</strong> B.t. on the leaf surface. These methods entail all aspects <strong>of</strong><br />

application technology: generating optimum-sized droplets that contain the optimum


226 M. A. Hulme and G. W. Green<br />

production <strong>of</strong> biological control agents is seen as a less attractive commercial proposition,<br />

and it is thus important that government organizations maintain an interest in providing<br />

opportunities for large-scale production <strong>of</strong> biological control agents.<br />

Economic aspects <strong>of</strong> various biological control strategies have been briefly mentioned<br />

in this chapter. These aspects need addressing in more detail to supplement the few cost<br />

estimates developed here by biologists. Analysis by economists could provide a more<br />

comprehensive picture <strong>of</strong> costs and benefits: cost figures should separate production<br />

from application; and benefit figures should include carry-over effects following the<br />

year <strong>of</strong> application.<br />

Target pests for control<br />

Judging by progress reported in this and other reviews <strong>of</strong> biological control there seems<br />

little reason to rule out any forest pest on a priori grounds as a suitable candidate for<br />

biological control. Each decision to attempt control must, <strong>of</strong> course, take into account<br />

the specific situation <strong>of</strong> pest and host tree(s) and the range <strong>of</strong> control options that might<br />

be applied. However, in considering biological control there seems good reason to give<br />

more attention to research opportunities with the key economic pests in <strong>Canada</strong>'s<br />

forests, which would include C. fumiferana in eastern <strong>Canada</strong> and D. ponderosae and D.<br />

ru/ipennis (Kby.) in western <strong>Canada</strong><br />

In the case <strong>of</strong> C. fumiferana, a good short-term option using biological control is now<br />

available commercially. Longer term control is desirable but biological control attempts<br />

have met with limited success. There is, <strong>of</strong> course, scope for further work and, although<br />

all attempts will be difficult, we suggest that searches for more virulent viruses and for<br />

exotic parasitoids be given special emphasis. Candidate viruses to date exhibit marginal<br />

virulence against C. fumiferana but some control has nevertheless been obtained and<br />

further work is warranted. Parasitoid introductions as a biological control option have<br />

been given little more than a cursory examination, perhaps because C. fumiferana is a<br />

native pest. Many <strong>of</strong> the releases were conducted in unsuitable conditions and a more<br />

systematic study <strong>of</strong> this control option would allow a more factual comparison <strong>of</strong> its<br />

potential with other control options. The most common candidate hosts in Europe that<br />

have been suggested for parasitoid collections, based on the similarity <strong>of</strong> their life cycles<br />

to C. fumiferana, are C. murinana Hb. and Epinolia nigricana L. on Abies spp., Dichelia<br />

histrionana Froel. on Picea spp., and Archips oporana L. on Larix spp., Pinus spp., and<br />

Picea spp. Several Zeiraphera spp. <strong>of</strong>fer more remote possibilities. Collection activities<br />

should not, however, be confined to Europe. China, for example, has a Choristoneura sp.<br />

that attacks Picea spp. and appears to be well controlled by natural factors (Macdonald &<br />

Pollard 1976).<br />

With Dendroctonus spp. <strong>of</strong> bark beetles, particularly D. ponderosae and D. ru/ipennis<br />

that are major pests in western <strong>Canada</strong> perhaps the whole range <strong>of</strong> biological control<br />

agents needs closer examination to see which, if any, have practical potential for control.<br />

Certainly a wealth <strong>of</strong> information exists on related scolytids such as the southern pine<br />

bark beetle, D. frontalis Zimmerman (Thatcher et al. 1978), some <strong>of</strong> which could provide<br />

a useful starting point in considering options with other Dendroctonus spp. More field<br />

examinations <strong>of</strong> the natural enemy complex <strong>of</strong> Dendroctonus spp. in <strong>Canada</strong> is obviously<br />

required. Multiphagous parasitoids <strong>of</strong> bark beetles that attack hardwoods should not be<br />

overlooked, as some also parasitize bark beetles that attack s<strong>of</strong>twoods.<br />

Other particularly destructive pests that are not included in this review are the<br />

hemlock looper, L. [lScellaria[lScellaria, and the western hemlock looper, L. [lScellaria<br />

lugubrosa (Hulst); populations are now at endemic levels and irruptions are likely within<br />

the next few years. For the present, however, the most logical approach is to give priority<br />

to biological control research, mainly in accordance with the current importance <strong>of</strong> the<br />

pest, as the scope <strong>of</strong> research opportunities is limited not by lack <strong>of</strong> suitable problems but<br />

by the resources we are able to devote to biological control. There seems little doubt that


Literature Cited<br />

Biological control <strong>of</strong> forest insect pests 227<br />

interest in this method <strong>of</strong> pest control is growing and more resources are likely to be<br />

committed in the future. This increased effort should in turn lead to progress that better<br />

reflects the potential <strong>of</strong> biological control as a viable pest management option.<br />

Beglyarov. G.A.; Smetnik. A.J. (1977) Seasonal colonization <strong>of</strong> entomophages in the USSR. In: Ridgway. R.L.; Vinson. S.B. (Eels.)<br />

Biological control by augmentation <strong>of</strong> natural enemies. New York: Plenum. pp. 283-328.<br />

Beirne. B.P. (1975) Biological control attempts by introductions against pest insects in the field in <strong>Canada</strong>. Canadian Entomologist 107.<br />

225-236.<br />

Bennett. F.D. (1974) Criteria for determination <strong>of</strong> candidate hosts and for selection <strong>of</strong> biotic agents. In: Maxwell. F.G.: Harris. F.A. (Eels.)<br />

Proceedings <strong>of</strong> a Summer Institute on Biological Control <strong>of</strong> Plant Insects and Diseases. Jackson.<br />

Mississippi; University Press <strong>of</strong> Mississippi. pp. 87-96.<br />

Bustillo. A.E.; Drooz. A.T. (1977) Comparative establishment <strong>of</strong> a Virginia (USA) strain <strong>of</strong> Telenomus alsophilae on Oxydia try'chiata in<br />

Columbia. Journal <strong>of</strong> Economic Entomology 70.767-770.<br />

Caltagirone, L.E. (1981) Landmark examples <strong>of</strong> classical biological control. Annual Review <strong>of</strong> Entomology 26. 213-232.<br />

DeBach. P.; Huffaker. C.B. : MacPhee. A. W. (1976) Evaluation <strong>of</strong> the impact <strong>of</strong> natural enemies. In: Huffaker, C. B.; Messenger, P .S. (Eels.)<br />

TIleory and practice <strong>of</strong> biological control. London, New York; Academic Press. pp. 255-285.<br />

Dorais. L.; Pelletier. M.: Smirn<strong>of</strong>f. W.A. (1980) Pulvcrisations acriennes de Bacillus thuringiensis Berliner realisees au Quebec de 1971 1\<br />

1979 contre la tordeusc des bourgeons de I·cpinette. Choristoneurafumiferana (Oem.). [Abstract) Rapport<br />

presentc au VI' Congres International de I'Aviation Agricole a Turin. Italic 22-26 Sept. 1980. unpaginated.<br />

Dulmage, H.T. (1970) Insectidical activity <strong>of</strong> HD-l. a new isolate <strong>of</strong> Bacillus thuringiensis val. alesti. Journal <strong>of</strong> Invertebrate Pathology 15.<br />

232-239.<br />

Gross. H. R. Jr. (1981) Employment <strong>of</strong> kairomones in the management <strong>of</strong> parasitoids. In: Norlund. D.A.: Jones. R.L.; Lewis. W.J. (Eels.)<br />

Semiochemicals and their role in pest control. New York; John Wiley. pp. 137-152.<br />

Huffaker. C.B.; Messenger. P.S.; DeBach. P. (1971) The natural enemy component in natural control and the theory <strong>of</strong><br />

biological control. In: Huffakcr. C. B. (Ed. ) Biological control. New York. London; Plenum. pp. 253-293<br />

Hussey. N.W. (1977) Biological control techniques. In: Rishbeth. R.S. (leader). The Royal <strong>Society</strong> delegation on<br />

biological control to China. Aug. 15 - Sept. 3. London; Royal <strong>Society</strong>. 38 pp.<br />

Knipling. E.F. (1979) The basic principles <strong>of</strong> insect population. suppression. and management. US Department <strong>of</strong> Agriculture Handbook<br />

512.659 pp.<br />

Macdonald. D.R.: Pollard. D.F.W. (1976) Report on the scientific exchange in forestry with the People's Republic <strong>of</strong> China. Canadian<br />

Forestry Service Report Pacific Forestry Research Centre. 32 pp.<br />

McFadden. M.W.; Dahlsten. D.L.; Berisford. C.W.; Knight. F.B.: Metterhousc. M.W. (1981) Integrated pest management in China's<br />

forests. Journal <strong>of</strong> Forestry 79. 723-726.<br />

McGugan. B.M.; Coppel. H.C. (1962) Biological control <strong>of</strong> forest insects 1910-1958. In: A review <strong>of</strong> the biological control attempts against<br />

insects and weeds in <strong>Canada</strong>. Commonwealth Institute <strong>of</strong> Biological Control Technical Communication<br />

2.35-127.<br />

Mackauer. M. (1976) Genetic problems in the production <strong>of</strong> biological control agents. Annual Rel'iew <strong>of</strong> Entomology 21.369-385.<br />

Morris. O.N. (1980) Entomopathogenic viruses: strategies for use in for usc in forest insect pest management. Canadian Entomologist 112.<br />

573-584.<br />

Munroc. E.G. (1971) Status and potential <strong>of</strong> biological control in <strong>Canada</strong>. In: Biological control programme against insects and weeds in<br />

<strong>Canada</strong>. Commonwealth Institute <strong>of</strong> Biological Control Technical Communication 4. 213-255.<br />

Otvos. I. (1973) Biological control agents and their role in the population fluctuation <strong>of</strong> the eastern hemlock looper in<br />

Newfoundland.Canadian Forestry Sen'ice Information Report N-X-102. 34 pp.<br />

Pimentel. D. (1963) Introducing parasites and predators to control native pests. Canadian Entomologist 95.785-792.<br />

Poinar. G.O. Jr. (1979) Nematodes for biological control <strong>of</strong> insects. Boca Raton. Florida: Chemical Rubber Company. 220 pp.<br />

Pschorn-Walcher. H. (1977) Biological control <strong>of</strong> forest insects. Annual Review <strong>of</strong> Entomology 22. 1-22.<br />

Reeks. W .A.; Cameron. J.M. (1971) Current approaches to biological control <strong>of</strong> forest insects 1959-1968. In: Biological control programmes<br />

against insects and weeds in <strong>Canada</strong>. Commonwealth Institute <strong>of</strong> Biological Control Technical Communication<br />

4. 105-112.<br />

Schroeder. D. (1974) A study <strong>of</strong> the interaction between internal larval parasites <strong>of</strong> Rhyacionia buoliana. Entomophaga 19. 145-171.<br />

Simmonds. F.J. (1969) Brief resume <strong>of</strong> activities and rccent successes achieved. Commonwealth Institute <strong>of</strong> Biological Control Report.<br />

Trinidad. 16 pp.<br />

Simmonds. F.J. (1972) Approaches to biological control problems. EnlOmophaga 17.251-264.<br />

Thatcher. R.C.; Searcy. J.L.; Coster. J.E.; Hertel. G.D. (1978) The southern pine beetle. US Department <strong>of</strong> Agriculture Technical Bulletin<br />

1631, 266 pp.<br />

Tropin.I.V.; Vedernikov. N.M.: Kranguaz. R.A.; Maslov. A.D.; Zubov. P.A.: Khramtsov. N.N.;Andreeva.G.I.;Lyashenko, L.1. (1980) A<br />

manual on the protection <strong>of</strong> the forest against pests and diseases. Moscow: Lcanaya promyshelnnost. 500 pp.<br />

Vinson. S.B. (1977) Behavioural chcmicals in the augmentation <strong>of</strong> natural cnemies. In: Ridgway. R.L.: Vinson. S.B. (Eels.) Biological<br />

control by augmentation <strong>of</strong> natural enemies. New York; Plenum. pp. 237-265.<br />

Weseloh. R.M. (1981) Host location by parasitoids. In: Nordlund. D.A.; Jones. R.L.: Lewis. W.J. (Eels.) Semiochemicals and their role in<br />

pest control. New York: John Wiley. pp. 79-96.


Blank Page<br />

228


Pest Status<br />

Chapter 46<br />

Background and Evaluation <strong>of</strong> Control Attempts<br />

Adelges piceae (Ratz.), Balsam<br />

Woolly Adelgid (Homoptera: Adelgidae)<br />

H.O. SCHOOLEY, l.W.E. HARRIS and B. PENDREL<br />

The balsam woolly adelgid, AdeJges piceae (Ratz.), is a destructive pest <strong>of</strong> firs, Abies<br />

spp. In recent years, however, populations have been low and dispersal has been<br />

limited. The distribution <strong>of</strong> the adelgid remains about the same as in 1968 (Clark el al.<br />

1971). In eastern <strong>Canada</strong>, expansion <strong>of</strong> the infestation has been largely into those small<br />

areas that had been bypassed as the initial infestation developed (Schooley 1980). The<br />

adelgid is now present throughout most <strong>of</strong> the fir forests <strong>of</strong> the Maritime Provinces,<br />

Newfoundland, and the eastern tip <strong>of</strong> the Gaspe Peninsula. A total area <strong>of</strong> about 103600<br />

kml is affected. In western <strong>Canada</strong> there have been only minor increases in the size <strong>of</strong><br />

the infestation, which now covers about 10 360 kml in southwestern British Columbia.<br />

The infestations in <strong>Canada</strong> extend southward into the United States on both the east and<br />

west coasts (Mitchell el al. 1970). Climate seems to be the main factor limiting further<br />

spread <strong>of</strong> this insect in <strong>Canada</strong> (Greenbank 1970).<br />

The balsam woolly adelgid attacks all species <strong>of</strong> firs in <strong>Canada</strong> but some are more<br />

resistant to injury than others. In eastern <strong>Canada</strong>, mortality <strong>of</strong> balsam fir, Abies balsamea<br />

(L.) Mill., has been common. In western <strong>Canada</strong>, subalpine fir, A. Jasiocarpa (Hook.)<br />

Nutt., growing at low elevations, is the most vulnerable species, but it rarely occurs<br />

within the present infestation boundaries (Harris 1968). Pacific silver or amabilis fir, A.<br />

amabilis (Doug!.) Forb., is the second most vulnerable western host and considerable<br />

mortality <strong>of</strong> this species has occurred. Grand fir, A. grandis (Doug!.) Lind!', is also<br />

deformed and killed by the adelgid, primarily on southern Vancouver Island. The<br />

adelgid is also threatening to destroy Fraser fir, A. fraseri (Pursh.) Poir., in the few<br />

remaining spruce-fir stands <strong>of</strong> the southern Appalachian Mountains <strong>of</strong> eastern United<br />

States (Johnson 1980).<br />

Some <strong>of</strong> the causes <strong>of</strong> variation in the response to adelgid infestation among Abies<br />

host species have been identified. Puritch (1973), for example, has shown that the onset<br />

and intensity <strong>of</strong> adelgid attack caused premature heartwood formation and a resulting<br />

abnormal water stress. Thus, ability to withstand this stress among fir species corresponds<br />

to their susceptibility to adelgid damage. It has also been observed that adelgid attack<br />

has little effect on European species because a layer <strong>of</strong> dead bark tissue develops during<br />

the wound healing process and protects the trees from further attack for several years<br />

(Oechssler 1962). This healing process is delayed or absent in balsam fir that is under<br />

continuous adelgid attack.<br />

Clark et al. (1971) stated that there was little scope for additional study in the control <strong>of</strong><br />

this adelgid by introduced predators. Field trials in eastern <strong>Canada</strong> tested all the<br />

apparently suitable insects over a 35-year period and failed to find a predator or predator<br />

complex that would control the adelgid. Only eight species <strong>of</strong> introduced predators have<br />

become established, at least temporarily, but none has been effective (Table 55) in<br />

reducing damage to economically tolerable levels. Consequently, it was recommended<br />

that the importation and release <strong>of</strong> predators be discontinued. This recommendation has<br />

229


230 H. O. Schooley. J.W.E. Harris and B. Pendrel<br />

been accepted and no introductions or releases have been made since 1969. The search<br />

for surviving introduced predators was discontinued in 1969, 1971. and 1978 in both New<br />

Brunswick and Nova Scotia (D.O. Greenbank. personal communication). in Newfoundland<br />

(Bryant 1971). and in British Columbia (Harris & Dawson 1979). The history <strong>of</strong><br />

predator releases and recoveries tabulated by Clark et al. (1971) has been updated<br />

Table 55 Open releases and recoveries <strong>of</strong> predators against Adelges piceae (Ratz.)<br />

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

Species and province Year Origin Number Recovery<br />

Adalia luteopicta Mulsant<br />

Newfoundland 1960 India 159<br />

Adalia ronina (Lewis)<br />

Newfoundland 1961 Japan 67<br />

Nova Scotia 1963 Japan 290<br />

New Brunswick 1960 Japan 25 1960<br />

1962 Japan 37 1962<br />

1963 Japan 585 1963<br />

Adalia tetraspilota (Hope)<br />

Newfoundland 1960 India 33<br />

Aphidecta obliterata (L.)<br />

Newfoundland 1959 Czechoslovakia 735<br />

1960 Germany 934<br />

1962 Czechoslovakia 848<br />

1963 Germany 989<br />

1964 Germany 1 187<br />

1965 Czechoslovakia 185<br />

Germany 267<br />

1966 Czechoslovakia 509<br />

1966 Austria 1380 1966<br />

1967 Austria 4288 1967<br />

1968 Germany 1096 1968<br />

and Austria<br />

Nova Scotia 1964 Germany 1 773 1967<br />

1966 Austria 370<br />

Norway 47<br />

1967 Norway 422<br />

New Brunswick 1962 Germany 1827 1962<br />

1963 Germany 3157 1963<br />

British Columbia 1960 Germany 1050<br />

1961 Germany 1 141<br />

1962 Germany 796<br />

1963 Germany 1997<br />

1965 Czechoslovakia 660<br />

1968 Germany 1069<br />

1969 Germany 420<br />

Aphidoletes thompsoni (Mohn)<br />

Newfoundland 1959 Czechoslovakia 3 124<br />

Germany 25 952<br />

1962 Germany 270 1961<br />

1963 Germany 5201


Tablc 55 continucd<br />

Adelges piceadRatz.), 231<br />

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

Species and province Year Origin Number recovery<br />

1965 Germany 826 1967<br />

1966 Germany 35119<br />

1968 Germany 7248<br />

Nova Scotia 1965 Germany 450 1965<br />

1966 Germany 37 110 1966-67<br />

New Brunswick 1959 Germany 36131 1959-61<br />

1965 Germany 551 1965<br />

1966 Germany 6300 1966<br />

British Columbia 1962 Germany 280<br />

1963 Germany 516<br />

1965 Germany 1080 1967<br />

1966 Germany 6845<br />

Balaustium sp.<br />

Quebec 1967 Pakistan 126<br />

Ballia eue/,aris Mulsant<br />

Newfoundland 1960 India 32<br />

New Brunswick 1959 Pakistan 79<br />

1960 India 55<br />

Coccinella septempunctata L.<br />

New Brunswick 1959 India 28<br />

1960 India 22<br />

Cremifania nigrocel/ulata Czerney<br />

Newfoundland 1959 Germany 198 1971<br />

1961 Germany 17<br />

Nova Scotia 1966 Germany 169 1966<br />

New Brunswick 1963 Germany 48 1964-65<br />

1966 Germany 17 1966-68<br />

British Columbia 1966 Germany 137<br />

1968 Germany 706 1969<br />

Exochomus lituratus Gorham<br />

Newfoundland 1960 Pakistan no<br />

Nova Scotia 1963 Pakistan 991<br />

New Brunswick 1963 Pakistan 209<br />

Exochomus uropygialis Mulsan!<br />

Newfoundland 1960 India 226<br />

1960 Pakistan 2839<br />

Nova Scotia 1963 Pakistan 8782<br />

New Brunswick 1959 Pakistan 2550<br />

1963 Pakistan 247<br />

1964 Pakistan 238<br />

Harmonia breiti Mader<br />

Newfoundland 1960 Pakistan 88<br />

New Brunswick 1959 India 85<br />

Laricobius erichsonii Rosenhauer<br />

Newfoundland 1959 Czechoslovakia 1043 1959<br />

Germany 2159 1960


232 H. O. Schooley. J.W.E. Harris and B. Pendrel<br />

Tahlc 55 continucd<br />

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

Species and province Year Origin Number recovery<br />

1960 Germany 8228 1961<br />

1961 Germany 1 118 1963<br />

1962 Germany 7940 1964<br />

1963 Germany 1869 1965<br />

Newfoundland<br />

via Europe 1984 1966<br />

1964 Germany 1819 1968<br />

Newfoundland<br />

via Europe 300<br />

1965 Germany 152<br />

Nova Scotia 1966 Germany 2575 1967<br />

New Brunswick 1960 Germany 14266 1958-68<br />

1966 Germany 3497 1958-68<br />

British Columbia 1960 Germany 800 1962-65<br />

1961 Germany 1432 1978<br />

1963 Germany 4871 1966,74<br />

1965 Germany 612<br />

1968 Germany 3164 1969,71,74<br />

Leucopis (Leucopis) n. sp. or.<br />

'melanopus'Tanasijtshuk*<br />

Newfoundland 1959 Germany 160 1959-62<br />

1968 Germany 1040<br />

New Brunswick 1960 Germany 166 1958-68<br />

1966 Germany 241<br />

British Columbia 1968 Germany 2273<br />

Leucopis (Neoleucopis)<br />

obscura Holiday<br />

Newfoundland 1965 Austria 24 1959-68<br />

Leucopis (Neolellcopis)<br />

alralula Ratz.<br />

New Brunswick 1965 Germany 385<br />

Pul/us impexus (Mulsant)<br />

Newfoundland 1959 Germany 9500<br />

1960 Germany 1 146 1960<br />

1961 Germany 131 1961<br />

1966 Germany 18036<br />

Nova Scotia 1963 Germany 1000 1965-67<br />

1964 Germany 2031 1965-67<br />

1966 Germany 4700 1965-67<br />

New Brunswick 1962 Germany 685 1958-59<br />

1963 Germany 397 1958-59<br />

1964 Germany 200 1958-59<br />

1966 Germany 26550 1958-59<br />

British Columbia 1960 Germany 1240 1961<br />

1963 Germany 1400<br />

1965 Germany 2417<br />

1966 Germany 18513<br />

1968 Germany 2079 1969,71.74.78


Table 55<br />

continued<br />

Aclelge.\· ph-elle (Ratz.). 233<br />

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

Species and province Year Origin Number recovery<br />

Scymnus pumilio (Weise)<br />

Newfoundland 1960 Australia 9687<br />

New Brunswick 1959 Australia 5590<br />

1960 Australia 7286<br />

British Columbia 1960 Australia 2930<br />

Telraphleps abdulghani Ghauri<br />

Nova Scotia 1965 Pakistan 949<br />

New Brunswick 1962 Pakistan 1972<br />

1963 Pakistan I 157<br />

1964 Pakistan 45<br />

1965 India 278<br />

Pakistan 2921<br />

British Columbia 1965 India 19<br />

Pakistan 1257<br />

Tetraphleps raoi Ghauri<br />

Nova Scotia 1965 India 59<br />

New Brunswick 1965 India 59<br />

• Species 'M'.<br />

Attempts to control the adelgid with contact or systemic insecticides have been<br />

unsuccessful. Several insecticides tested proved suitable for ornamental and nursery<br />

applications where it was possible to saturate infested trees with chemical solutions.<br />

However, not one <strong>of</strong> the 51 insecticides used in laboratory and greenhouse experiments<br />

caused acceptable levels <strong>of</strong> adelgid mortality under field conditions (Nigam 1972, 1976).<br />

Recently, some success has been obtained in controlling adelgids on individual trees<br />

with insecticidal soaps (Puritch 1975). These chemicals are expected to be useful only<br />

for ornamental and nursery applications.<br />

There are no known insect parasites <strong>of</strong> the balsam woolly adelgid but several fungal<br />

diseases have been reported. including Fllsarium larvaTllm Forbel and Cephalosporium<br />

coccoTllm Petch found in the Gaspe region <strong>of</strong> Quebec (Smirn<strong>of</strong>f 1970), Cephalosporium<br />

sp. and Penicillium sp. found in British Columbia (Harriset al. 1966), and F. nivale (Fries)<br />

Cesati found on adelgids in North Carolina. USA (Fedde 1971). Greenhouse and field<br />

experiments with the diseases from Quebec have been unsuccessful. The potential <strong>of</strong><br />

organisms from other locations as natural control agents remains unknown.<br />

Failure to control the balsam woolly adelgid is viewed with special concern in Atlantic<br />

<strong>Canada</strong>. In Nova Scotia and Prince Edward Island, balsam fir stands are usually<br />

damaged before reaching a marketable size. On the island <strong>of</strong> Newfoundland, spectacular<br />

and unprecedented mortality <strong>of</strong> balsam fir has occurred on an accelerated scale in stands<br />

where adelgid attack has been followed by severe spruce budworm defoliation (Schooley<br />

1981). In affected areas. stand conversion. i.e. the replacement <strong>of</strong> balsam fir with other<br />

tree species, is being conducted on a large scale. Black spruce, Picea mariana (Mill.)<br />

B.S.P., is the favoured alternate species but white spruce. P. glauca (Moench) Voss,<br />

eastern larch. Larix larici"a (Du Roi) K. Koch. and hybrid larch species are being<br />

considered.


234 H. O. Schooley. J. W. E. Harris and B. Pendrel<br />

Recommendations<br />

Literature Cited<br />

As previously recommended by Clark et al. (1971) no further introduction <strong>of</strong> predators<br />

should be made into eastern <strong>Canada</strong> until new species that might effectively control<br />

the adelgids are identified. Suitable predators may be found attacking two closely related<br />

fir shoot adelgids. Adelges merkeris Eichhorn and Adelges nusslini C.B .• in Europe. The<br />

milder winter temperatures <strong>of</strong> coastal British Columbia could favour a species that failed<br />

in the east. Also. some species already tested on the west coast may have failed because<br />

too few individuals were released or because <strong>of</strong> unfavourable weather conditions at the<br />

time <strong>of</strong> release. Therefore. further introductions may be considered for western <strong>Canada</strong>.<br />

Any predator successfully introduced and proved to be effective in western <strong>Canada</strong><br />

should be considered for testing in the mild coastal areas <strong>of</strong> eastern <strong>Canada</strong>.<br />

Population dynamics studies should be conducted regularly in all regions where the<br />

adelgid is present. These studies should include observations that will identify any<br />

factors that successfully control mortality. Financial support should also be made<br />

available for periodic reappraisal <strong>of</strong> the effect that natural insect and disease organisms<br />

have in controlling the balsam woolly adelgid in other countries.<br />

Bryant. D.G. (1971) Forest protection: balsam woolly aphid. Adelges piceae (Ratz.). Annual Report Newfoundland Forest Protection<br />

Associalion 1971. 43 PI'.<br />

Clark, R.e.; Greenbank, D.O.; Bryant. D.G.; Harris, J.W.E. (1971) Adelges piceae (Ratz.). balsam woolly aphid (Homoptera: Adclgidae).<br />

In: Biological control programmes against insects and weeds in <strong>Canada</strong>. Commonwealth Institute <strong>of</strong><br />

Biological Control Technical Communication 4. 113-127.<br />

Fedde. G.F. (1971) A parasitic fungus disease <strong>of</strong> Adelges piceae (Homoptera: Phylloxeridae) in North Carolina. Annals <strong>of</strong> the <strong>Entomological</strong><br />

<strong>Society</strong> <strong>of</strong> America 64, 749-750<br />

Greenbank, D.O. (1970) Climate and the ecology <strong>of</strong> the balsam woolly aphid. Canadian Entomologist 102. 546-578.<br />

Harris. J.W.E. (1968) Balsam woolly aphid in British Columbia. Canadian Department <strong>of</strong> Forestry and Rural Development Forestry<br />

Researcl. Forest Pest Leaflet B.C.-I, 5 PI'.<br />

Harris, J. W. E.; Dawson. A. F. (1979) Predator release program for balsam woolly aphid, Adelges piceae (Homoptera: Adelgidae), in British<br />

Columbia, 1960-1969. Journal <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> British Columbia 76, 21-26.<br />

Harris. J.W.E.; Allen. S.1.; Collis, D.G.; Harvey, E.G. (1966) Studies <strong>of</strong> the balsam woolly aphid. Adelges piceae (Ratz.), in British<br />

Columbia. Canadian Department <strong>of</strong> Forestry Information Report BC-X-5, 5 pp.<br />

Johnson, K. (1980) Fraser fir and balsam woolly aphid. Summary <strong>of</strong> information. Southern Appalachian Reserve Resources Management<br />

Commillee Report, 62 pp.<br />

Mitchell. R.G.; Amman, D.G.; Waters. W.E. (1970) Balsam woolly aphid. US Department <strong>of</strong> Agriculture Forestry Service, Forest Pest<br />

Leaflet 118. 10 PI'.<br />

Nigam. P.e. (1972) Summary <strong>of</strong> toxicity <strong>of</strong> insecticides and chemical control studies against balsam woolly aphid. Canadian Forestry Sen'ice<br />

Information Report CC-X-26. 7 pp.<br />

Nigam. P.C. (1976) Summary <strong>of</strong> chemical control studies against balsam woolly aphid in British Columbia. 1973-74.Canadian Forestry<br />

Service Information Report CC-X-133, 6 PI'.<br />

Oechssler. G. (1962) Sucking injuries to the tissue <strong>of</strong> native and exotic firs caused by central European fir aphids. Zei/Schriftfiir angl'wandte<br />

EnlOmologie SO, 418-459 (Canadian Department <strong>of</strong> Forestry Translation 57).<br />

Puritch. G.S. (1973) The effects <strong>of</strong> water stress on photosynthesis. respiration and transpiration <strong>of</strong> four Abies species. Canadian Journal <strong>of</strong><br />

Forl'st Rl'search 3. 293-298.<br />

Puritch. G.S. (1975) The toxic effects <strong>of</strong> fatty acids and their salts on the balsam woolly aphid. Adelges piceae (RalZ.). Canadian Journal <strong>of</strong><br />

Forest Resl'arch 5, 515-522.<br />

Schooley. H.O. (1980) The introduction. spread and occasional resurgence <strong>of</strong>the balsam woolly aphid in Newfoundland. Proceedings <strong>of</strong> thl'<br />

International Union Forest Research Organizations Conference on Dispersal <strong>of</strong> Forest Insec/S, Aug.Lrt<br />

27-31. 1979. Washington State University, Pullman, Washington. PI'. 116-127.<br />

Schooley, H.O. (1981) An evaluation <strong>of</strong> the hazard rating systcm for balsam woolly aphid damage in Newfoundland. Procel'dings <strong>of</strong> a<br />

Symposium on Hazard Rating Systems in Forest Insect Pest Management, July 31 - August I, 1980.<br />

Athens, Georgia. US Department <strong>of</strong> AgriculTUre Forestry Service General Technical Report WO-27.<br />

Smirn<strong>of</strong>f. W.A. (1970) Fungus diseases affecting Adelges piceae in the fir forest <strong>of</strong> the Gaspe Peninsula. Quebec. Canadian Entomologist<br />

102. 799-805.


Pest Status<br />

Chapter 47<br />

Choristoneura fumiferana (Clemens),<br />

Spruce Budworm (Lepidoptera:<br />

Tortricidae)<br />

C.A. MILLER<br />

The spruce budworm, CllOristoneura fumiferana (Clem.). is a native defoliator <strong>of</strong> the<br />

spruce, Picea spp., and balsam fir. Abies balsamea (L.) Mill., forests <strong>of</strong> North America,<br />

which extend from northern Alberta to the Atlantic seaboard. It is the most important<br />

economic pest in forests <strong>of</strong> eastern <strong>Canada</strong> because <strong>of</strong> the frequency <strong>of</strong> epidemics and<br />

the intensity and extent <strong>of</strong> fibre losses. Budworm populations periodically irrupt to<br />

epidemic proportions and cause extensive tree mortality and reduced growth in surviving<br />

trees. There have been three major epidemics in this century. The first occurred around<br />

1915 and covered most <strong>of</strong> the eastern Quebec - Atlantic Region; the second irrupted in<br />

Ontario - western Quebec in the late 1930s and appeared in the Atlantic Region in the<br />

late 1940s, but had little impact in Nova Scotia and Newfoundland; and the current<br />

outbreak irrupted in the late 1960s. In sequence, these three outbreaks <strong>of</strong> the 20th<br />

century have tended to increase in size and frequency.<br />

The review period <strong>of</strong> this publication, 1969-80 largely coincides with the third major<br />

epidemic. The prolonged second epidemic collapsed in the 1960s, except for a persistent<br />

infestation covering less than a million hectares in central New Brunswick. During the 4<br />

years 1967 - 70, spruce budworm population increases were recorded in all provinces<br />

from Ontario to Newfoundland, and by 1975 the epidemic reached its peak, covering<br />

about 54 x 10 6 ha, or about half <strong>of</strong> the land area in eastern <strong>Canada</strong> classified as productive<br />

forest. Much <strong>of</strong> the remainder is not occupied by spruce and fir.<br />

In general, the intensity <strong>of</strong> the infestation, as measured by egg-mass densities across<br />

eastern <strong>Canada</strong>, declined in the latter part <strong>of</strong> the 1970s but the density <strong>of</strong> feeding larvae<br />

relative to the decreased quality <strong>of</strong> foliage on previously attacked trees remained high<br />

enough to cause severe damage. Thus in 1980 the epidemic still caused moderate to<br />

severe defoliation over 36 x 10" ha and it is expected that extensive defoliation will<br />

continue into the mid-1980s.<br />

Maturing balsam fir trees in unprotected forests usually die after 3 or 4 successive<br />

years <strong>of</strong> severe spruce budworm attack. Spruce trees, particularly black spruce, Picea<br />

mariana (Mill.) B.S.P., are more tolerant <strong>of</strong> defoliation and may withstand up to 6 years<br />

<strong>of</strong> attack. In the current outbreak, tree mortality was first recorded in northeastern<br />

Ontario and western Quebec in 1972, about 5 years after the beginning <strong>of</strong> the epidemic.<br />

In Nova Scotia (Cape Breton Highlands) and Newfoundland, mortality was first recorded<br />

between 1975 and 1976. Throughout the whole region, the area <strong>of</strong> vulnerable forest<br />

containing dying and dead trees increased each year and by 1980 was estimated at about<br />

19 x 10" ha (Table 56). Although 'zones <strong>of</strong> heavy defoliation' and 'zones <strong>of</strong> tree mortality'<br />

were delineated, it has proved difficult to translate these spatial data to a volumetric<br />

estimate <strong>of</strong> fibre loss. Inventory data on spruce and fir before the outbreak are meagre;<br />

moreover, the rapid detection <strong>of</strong> dying stands <strong>of</strong> trees by remote sensing is not yet<br />

operational. However, some estimates <strong>of</strong> fibre loss in mortality zones are available<br />

(Table 56). Generally, in eastern <strong>Canada</strong> in the late 1970s there was (1) continuing loss <strong>of</strong><br />

radial and height growth in severely defoliated stands; (2) tree mortality ranging from 10<br />

to 80% at the stand level in a mortality zone <strong>of</strong> about 19 x 10" ha; (3) a loss in fibre quality<br />

and volume due to secondary pest attack in dead trees that were salvaged; and (4)<br />

235


236 C. A. Miller<br />

Table 56<br />

Table 57<br />

evidence that the future marketable volume <strong>of</strong> young stands that recover from the epidemic<br />

may be significantly lower than the site potential (Baskerville & Maclean 1979).<br />

Estimates <strong>of</strong> tree mortality in eastern <strong>Canada</strong> during the spruce budworm, Clwristoneura<br />

fumiferana (Clem.), outbreak <strong>of</strong> the 1970s·<br />

Dead and dying<br />

trees, 1980<br />

Date <strong>of</strong> first Area Volume<br />

Province tree mortality (x 10" hal (x 10" ml)<br />

Ontario 1972 8.4 ?<br />

Quebec 1972 9.3 ?<br />

New Brunswick •• 0.7 30<br />

Nova Scotia 1976 16<br />

Newfoundland 1975 0.4 40<br />

• Data extracted from: Forest Insect and Disease Conditions in <strong>Canada</strong> 1980. Canadian<br />

Forestry Service, Ottawa, 1981.<br />

•• Persistent infestation in central New Brunswick from the mid-1950s, with some tree<br />

mortality.<br />

No comprehensive policy was developed to deal with the spruce budworm problem in<br />

the 1970s in eastern <strong>Canada</strong>. The main reason is that each province administers its own<br />

forest resource and each gave a different priority to the social, economic, and political<br />

problems caused by spruce budworm. For example, in New Brunswick (Table 57) where<br />

the forest industry is <strong>of</strong> great importance to the provincial economy and the rate <strong>of</strong><br />

utilization <strong>of</strong> spruce and fir is high, the policy was to protect Crown forests and designated<br />

private holdings where there was risk <strong>of</strong> severe tree damage and imminent tree mortality<br />

(Miller & Kettela 1975). Extensive annual use <strong>of</strong> chemical insecticides was the principal<br />

tactic in implementing this policy. In contrast, Ontario adopted a policy limiting chemical<br />

S<strong>of</strong>twood utilization relative to spruce, Picea spp., and balsam fir, Abies balsamea (L.)<br />

Mill., supply in New Brunswick and Ontario<br />

Pre-outbreak 1973<br />

Average<br />

Potential annual area<br />

Estimated Estimated A verage s<strong>of</strong>twood··· wood <strong>of</strong><br />

spruce and fir annual harvest 1979 supply sprayed<br />

volume·<br />

(x 10' ml)<br />

growth··<br />

(x 10' ml)<br />

Allowable<br />

(x 10' ml)<br />

Actual<br />

(x 10' ml)<br />

(D-C) forest (ha)<br />

1975-79<br />

A<br />

New Brunswick 348<br />

Ontario 1744<br />

B<br />

8.7<br />

44····<br />

C D<br />

7.1 7.4 Deficit<br />

27.1 19.3 Excess····<br />

• Statistics <strong>Canada</strong>, 1973 .<br />

•• Broad-scale estimate .<br />

••• F.L.C. Reed and Associates ltd., Canadian Forestry Congress. 1980 .<br />

•••• No correction for accessibility.<br />

2.3 x 10"<br />

4.0 x 10'


Literature Cited<br />

ChorislOIU'lIfCI fllmiferallil (Clemens). 237<br />

control programmes to high investment areas such as parks, nurseries, and high value<br />

timber stands, because in that province wood supply problems are less pressing and fir is<br />

a minor component <strong>of</strong> utilized s<strong>of</strong>twood. In all provinces, the issue <strong>of</strong> protection kindled<br />

debate on the use <strong>of</strong> chemicals. Various citizen groups argued that the use <strong>of</strong> chemical<br />

insecticides in forest spraying created unacceptable environmental and human health<br />

risks; they called for biological control and other pest management options. All provincial<br />

administration reacted to these concerns, but in Nova Scotia and Newfoundland the<br />

reaction resulted in curtailment <strong>of</strong> chemical spraying in forests that, with no intervention,<br />

had already suffered severe defoliation and increasing mortality.<br />

Although there has been research into many potential spruce budworm control techniques,<br />

operational control tactics in the face <strong>of</strong> a widespread epidemic are largely limited to<br />

removal <strong>of</strong> vulnerable stands and the suppression <strong>of</strong> feeding larval populations with<br />

chemical insecticides or with Bacillus Ihuringiensis Berliner. The following sections<br />

summarize the progress made with each <strong>of</strong> the biological control methods.<br />

Baskerville. G.L.; MacLean. D.A. (1979) Budworm caused mortality and 20-year recovery in immature balsam fir stands. Canadian<br />

Foreslry Sen'ice, Marilime Foresl Research Cenlre Informalion Reporl M-X-102.<br />

Miller. CA.; Kettela. E.G. (1975) Aerial control operations against the spruce budwonn in New Brunswick. 1952-1973. In: Prebble M.L.<br />

(Ed.) Aerial control <strong>of</strong> forest insects in <strong>Canada</strong>. Ottawa. Ontario; Thorn Press. pp. 95-112.


23R W. A. Smirn<strong>of</strong>fand O. N. Morris<br />

Background<br />

A. Field Development <strong>of</strong> Bacillus thuringiensis Berliner<br />

in Eastern <strong>Canada</strong>, 1970-80<br />

W.A. SMIRNOFF and O.N. MORRIS<br />

Bacillus Ihuringiensis var. kurslaki, (B.I.) is a bacterium registered in <strong>Canada</strong> for control<br />

<strong>of</strong> agricultural and forest pests. The registered preparations, derived from cultures <strong>of</strong> the<br />

serotype 3a, 3b, are a mixture <strong>of</strong> dormant endospores and endotoxin crystals suspended<br />

in a liquid carrier.<br />

Forest protection agencies have found spraying to be more expensive with B.I. than<br />

with chemical insecticides, yet B.I. remains a desirable alternative because its toxicity is<br />

confined to lepidopterous larvae and its use rarely arouses public antagonism. The<br />

Canadian experience <strong>of</strong> B.I. aerially applied against spruce budworm, Chorisloneura<br />

fumiferana (Clem.) is that its efficacy is variable. Field trials have indicated differences<br />

among commercial products, varying efficacies among field formulations, uncertainties<br />

over dosages, technical difficulties in spraying methods, weaknesses in means <strong>of</strong> measuring<br />

deposits, and inadequacies in techniques <strong>of</strong> assessing the agent's effectiveness in reducing<br />

larval populations and protecting foliage.<br />

Research into these problems was conducted in the 1970s, in two main areas, by two<br />

laboratories <strong>of</strong> the Canadian Forestry Service (C.F.S.): (a) in Quebec, where the<br />

Laurentian Forest Research Centre collaborated with the Quebec Department <strong>of</strong> Lands<br />

and Forests, and (b) in Ontario and the Atlantic Provinces, where the Forest Pest<br />

Management Institute participated in provincial-federal field trials. This paper collates<br />

the results from both laboratories.<br />

Mode <strong>of</strong> action<br />

B. Ihuringiensis serotype 3a, 3b is a bacterium that is pathogenic to lepidopterous<br />

larvae. On spruce budworm and other tortricids the commercial formulations cause<br />

enterotoxicosis by the action <strong>of</strong> the crystal-like parasporal inclusions (= delta endotoxin),<br />

followed by septicaemia from the action <strong>of</strong> their spores (Smirn<strong>of</strong>f & Valero 1979, Fast<br />

1981). The mode <strong>of</strong> action <strong>of</strong> the crystal toxin is described by Cooksey (1971).<br />

The pathogenic action <strong>of</strong> 8.1. on spruce budworm depends on the dose ingested. the<br />

age and physiological state <strong>of</strong> the larvae, the temperature, and possibly on the presence<br />

<strong>of</strong> natural infections by microsporidia. The optimal temperature for this action is<br />

20-22°C (Smirn<strong>of</strong>f 1967). Young larvae are more susceptible to B.I. than fifth- and sixthinstar<br />

larvae (Morris 1973).<br />

Formulations<br />

Initially B.I. was tested against spruce budworm in New Brunswick in 1960. The<br />

material used was Thuricide® SO-75 (Bi<strong>of</strong>erm Corp.), based on the Berliner serotype.<br />

Next, a small-scale aerial application <strong>of</strong> Thuricide® 90T was made in 1969, using both<br />

suspensions in water and water-in-oil emulsion (Morris el al. 1975). Assessment <strong>of</strong><br />

population density reduction and foliage protection was inconclusive owing to clogging<br />

<strong>of</strong> the spray nozzles and poor distribution <strong>of</strong> the active ingredient.<br />

Improved commercial preparations <strong>of</strong> B.I. became available for testing between 1970<br />

and 1980, namely Thuricide® 24B, Thuricide® 32B, Dipel® 36B (Abbot Ltd.), and<br />

Novabac® 32B (Cyanamid <strong>Canada</strong>) or Novabac® wettable powder titrating 80 000


A. Field development <strong>of</strong> Bacillus Ihurillgiellsis 239<br />

I. U .Img. These concentrates included various additives designed to improve deposit<br />

efficiency and adherence to foliage (rain fastness), and to prolong residual activity <strong>of</strong> the<br />

biological agent.<br />

Three field formulations were developed at the Laurentian Forest Research Centre,<br />

and have been used experimentally in Quebec. The first formulation, based on an<br />

aqueous solution <strong>of</strong> sodium dihydrogen phosphate (NaH 2P04), and the second, based<br />

on an aqueous solution <strong>of</strong> sorbitol, incorporate chitinase and Chevron® sticker (Chevron<br />

Chemical (<strong>Canada</strong>) Ltd.), and have been used since 1973 in low-volume sprays<br />

(4.7 l/ha). Sorbitol is a humectant, whereas sodium dihydrogen phosphate is added to<br />

increase the specific gravity <strong>of</strong> the liquid B.t. formulation. The third formulation, named<br />

Futura by the Laurentian Forest Research Centre, also uses sorbitol, chitinase, and<br />

sticker but incorporates less aqueous carrier and has been used for reduced-volume<br />

sprays (2.5 Jlha) since 1978 (Smirn<strong>of</strong>f 1980a, 1981a). Formulations used in Ontario and<br />

the Atlantic Provinces were similar, but did not include chitinase, sorbitol, or sodium<br />

dihydrogen phosphate (Morris 1980).<br />

Smirn<strong>of</strong>f (1973, 1977) concluded after several years <strong>of</strong> experiments that the addition <strong>of</strong><br />

minute quantities <strong>of</strong> chitinase increases the efficacy <strong>of</strong> B.t. against spruce budworm; in<br />

his view, it enhances penetration <strong>of</strong> ingested ingredients through the gut lining, and<br />

rapidly potentiates symptoms <strong>of</strong> infection (lethargy, weight loss) even at low temperatures.<br />

Some other additives have been tested for compatibility. Smirn<strong>of</strong>f (1981b) found that<br />

either mineral oil in large quantities or various dyes inhibited spore development.<br />

During 1970-79, Thuricide® 16B (Sandoz. Inc.) was the only product fully registered<br />

for spruce budworm control under the Canadian Pest Control Products Act; around 1980<br />

Dipel® 88 and Novabac® (Cyanamid <strong>Canada</strong>) have been fully registered.<br />

Spray delivery systems and measurement <strong>of</strong> deposit<br />

Sprays were applied by various aircraft ranging from four-engined freight planes to<br />

single-engined biplanes and helicopters. They included the Lockheed Constellation L-<br />

749, Douglas DC-6B, <strong>Canada</strong>ir CL-215, Grumman Avenger TBM, Grumman Ag-cat,<br />

Boeing Stearman, Piper Pawnee, Cessna Agtruck, and the Sikorsky S55-T helicopter.<br />

Several kinds <strong>of</strong> spray emission hardware were tried, including Beecomist® (Beeco<br />

Products Co., USA), Micronair® (Micronair (Aerial) Ltd., U.K.) boom and flat fan<br />

Teejet® nozzles, and boom and open nozzles. Spray aircraft were calibrated to ensure<br />

appropriate emission rates for prescribed swath widths. Swath width varied from 305 m<br />

for a Constellation to 38 m for an Ag-cat (Smirn<strong>of</strong>f 1979a, Smirn<strong>of</strong>f & Juneau 1982).<br />

There is no general agreement on the optimal droplet size or optimal foliage coverage.<br />

Spray coverage is measured as the number <strong>of</strong> droplets per square centimetre on Kromekote®<br />

cards, or as the number <strong>of</strong> colonies per square centimetre collected on agar plates<br />

or MiIlipore® filters, placed at ground level in forest clearings. The number <strong>of</strong> spores<br />

deposited per unit area on glass plates at ground level has also been measured (Morris<br />

1980). In 1979, Smirn<strong>of</strong>f & Valero proposed a new method whereby a given volume <strong>of</strong><br />

peptonized water was exposed to the spray, so that the number <strong>of</strong> spores deposited per<br />

unit area could be calculated. For B.t. treatments, this value is more important than the<br />

total volume <strong>of</strong> liquid deposited (Smirn<strong>of</strong>f 1980a). Field tests have shown the necessity<br />

for calibration <strong>of</strong> the spray system: assessment based on the viable spore count deposited<br />

per unit area provides an adequate calibration.<br />

The concentration <strong>of</strong> B.t. within the proprietary product, the properties <strong>of</strong> the<br />

adjuvants added at the airstrip mixing plant, and the composition <strong>of</strong> the tank concentrate<br />

all affect the rate <strong>of</strong> deposition and the success <strong>of</strong> operations (Smirn<strong>of</strong>f 1980b).<br />

Operators used various kinds <strong>of</strong> spray hardware designed for chemical insecticides.<br />

Beecomist® spray heads were satisfactory on the Sikorsky helicopter. However, a


240 W. A. Smirn<strong>of</strong>f and O. N. Morris<br />

Field Trials<br />

boom and nozzle spray system composed <strong>of</strong> Teejet® 8004 flat fan nozzles gave better<br />

results with a Grumman Ag-cat fixed-wing aircraft (Smirn<strong>of</strong>f 1979a). A larger aircraft<br />

such as the DC-6B is used with open type nozzles, provided that the pumping system is<br />

able to produce a 413 kPa to 482 kPa pressure in the booms.<br />

In summary, a variety <strong>of</strong> means was used in the development <strong>of</strong> B.I. application<br />

technology: field trials using new formulations; improved hardware and nozzle systems;<br />

atomization and distribution <strong>of</strong> B.I.; knowledge <strong>of</strong> meteorological parameters; swath<br />

paths; improved deposit analysis; dosage rates; application timing and frequency;<br />

methods <strong>of</strong> measuring treatment effectiveness; and knowledge <strong>of</strong> the spruce budworm<br />

population densities most amenable to effective control (Morris 1980, Morris el al. 1981,<br />

Smirn<strong>of</strong>f 1979a, 1979b, 1980a, Auger el al. 1981).<br />

Assessment <strong>of</strong> efficacy<br />

The assessment <strong>of</strong> efficacy continues to be a major problem because both criteria -<br />

foliage saved and number <strong>of</strong> larvae killed - present sampling difficulties. The estimate<br />

<strong>of</strong> foliage saved by treatment is derived from comparisons <strong>of</strong> foliage persisting in<br />

untreated check plots with foliage retained in the sprayed plot. An important element <strong>of</strong><br />

the comparison is the production <strong>of</strong> new buds as a measure <strong>of</strong> foliage potential in the<br />

year following treatment. Quantification <strong>of</strong> foliage saved is based on counts <strong>of</strong> shoots,<br />

but the method lacks sensitivity because all shoots do not develop consistently. Since 1978<br />

an improved method has been developed at the Laurentian Forest Research Centre based<br />

on weight <strong>of</strong> new foliage produced and retained; it compares the weight <strong>of</strong> foliage<br />

produced by a defoliated tree with the expected weight <strong>of</strong> foliage produced by a healthy<br />

tree <strong>of</strong> similar branch area.<br />

The estimation <strong>of</strong> larval population reduction has usually been based on a comparison<br />

<strong>of</strong> the survival rate in the treated plot with that in an untreated check plot, using Abbott's<br />

formula to isolate the effect <strong>of</strong> the spray. Most workers recognize that this technique<br />

has major problems <strong>of</strong> sensitivity owing to the variability encountered in sampling.<br />

Non-lethal residual effects from B.I. have been observed by Morris (1973), Smirn<strong>of</strong>f<br />

(1978), and Morris el al. (1981); the influence <strong>of</strong> these effects on the dynamics <strong>of</strong>treated<br />

budworm populations cannot yet be assessed, but should not be ignored. Smirn<strong>of</strong>f<br />

detected changes in the "vitality" <strong>of</strong> survivors in treated populations, calculated by<br />

measurements <strong>of</strong> pupal weight, total lipids, calcium, and various enzymes. In a comparison<br />

<strong>of</strong> pupal populations in stands treated with the chemical insecticide fenitrothion,<br />

stands treated with B.I., and untreated stands, the survivors exposed to B.t. had the<br />

lowest vitality assessed by these biochemical criteria (Smim<strong>of</strong>f 1979c).<br />

Province <strong>of</strong> Quebec<br />

The field trials (Table 58) had various objectives, but were especially designed to test<br />

formulations and to measure deposits, larval mortality, and foliage protection under<br />

operational and experimental circumstances. The first trials in 1971 and 1972 encouraged<br />

the expectation that spruce budworm mortality and defoliation control could be achieved<br />

by adequate deposits <strong>of</strong> B.t. (Smim<strong>of</strong>f el al. 1973a, 1973b). Small trials in 1973 showed<br />

that B.t. sprays might be ineffective where spruce budworm populations were very high<br />

and stands were already weakened (Smirn<strong>of</strong>f el al. 1974). In 1974 and 1975, four-engined<br />

aircraft were used on a large scale for the first time, but the results were inconclusive<br />

because <strong>of</strong> improper calibration <strong>of</strong> the spray system (Smirn<strong>of</strong>f el al. 1976, Smirn<strong>of</strong>f &


244 W. A. Smirn<strong>of</strong>fand O. N. Morris<br />

Table 61<br />

Table 62<br />

Tests <strong>of</strong> Bacillus Ihuringiensis Berliner aerially applied against the spruce budworm,<br />

Clrorisloneura fumiferana (Oem.), on red spruce Picea rubens Sarg., in Nova Scotia and<br />

New Brunswick 1979-80"<br />

Ground<br />

Dosage deposit Pre·spray larval Percentage<br />

applied rate density per Percentage deroliation roliage<br />

Formulation"" Year (LU. x I0 9 iba) (drops/em l ) 45-cm branch Expected Actual saved<br />

Nova Scotia<br />

Thuricidcs 168 1979 I x 20 38-56 22 60 14 46<br />

Thuricidcs 168 1979 I x 20 51 14 38 9 28<br />

Thuricidc s 168 1979 2 x 20 27-31 12 36 14 22<br />

Thuricide!l 168 1900 I x 20 24 7 25 0.2 25<br />

Thuricidc:s 168 1900 1 x 20 30 7 25 14 II<br />

Thuricide!': 168 1900 I x 20 19 6 25 12 J3<br />

New Brunswick<br />

Dipell'l88 1979 1 x oW 20 17 46 23 23<br />

Thuricides 168 1979 I x 20 12-33 19 50 28 22<br />

" Includes only tests reporting pre-spray larval densities.<br />

"" Includes only formulations without chemical additives, e.g. chemical pesticides and<br />

enzymes.<br />

Extent <strong>of</strong> acceptable protection with Bacillus Ilruringiensis Berliner in eastern <strong>Canada</strong><br />

1979-80·<br />

Mean no. <strong>of</strong> Percentage <strong>of</strong><br />

Area droplets/em 2 area acceptably<br />

Province trea ted (ha) (range) protected""<br />

Ontario 4846 22 (11-30) 74<br />

- FPMI (1980 only) 320 34 (24-40) 100<br />

New Brunswick (1979 only) 360 23 (2-53) 100<br />

Nova Scotia 30188 51 (9-111) 92<br />

Newfoundland 15699 9 (5-16) 48<br />

and Labrador<br />

Quebec 33 174 43 (10-87) 89<br />

""<br />

Evaluation <strong>of</strong> Control Attempts<br />

Based on efficacy data provided by provincial departments or agencies.<br />

Acceptable protection is less than 50% defoliation <strong>of</strong> current year's growth.<br />

protected, ranged from 74 to 100% except for Newfoundland and Labrador, where the<br />

success rate was only 48%. The poor efficacy in Newfoundland was apparently due to<br />

inadequate spray deposits.<br />

The encouraging results with B.I. up to 1978 prompted the C.F.S. to formulate technical<br />

guidelines for the use <strong>of</strong> this product against the spruce budworm. A preliminary C.F.S.<br />

policy statement recommended "the operational testing <strong>of</strong> B.I. in environmentally<br />

sensitive areas or in sensitive areas where the forest manager either chooses not to use


Recommendations<br />

A. Field development <strong>of</strong> Bilcilllls IllIlringiensis 245<br />

or is forbidden to use conventional chemical insecticides for the control <strong>of</strong> the spruce<br />

budworm". The main technical recommendations were as follows:<br />

Objective - primarily to limit the average annual defoliation to less than 50% and<br />

secondarily to reduce the numbers <strong>of</strong> spruce budworm by 75-85%.<br />

Formulation - Thuricide® 16B. water (not chlorinated). Chevron® sticker (0.1 %).<br />

Optional for experimental trials: Thuricide lt 32B. sorbitol. Chevron® sticker. chitinase<br />

(9880 nephelometric units per hectare). water. and Erio Acid Red (experimental tracer).<br />

Note that dyes may inhibit B.I. efficacy.<br />

Dosage - 20 x 10" I.U. <strong>of</strong> B.I. in a minimum <strong>of</strong> 4.7 I/ha. In mixed spruce·fir stands. two<br />

applications <strong>of</strong> 10-20 x 10" I. U .Iha each.<br />

Operational constraints - pre·spray population density <strong>of</strong> 25-30 larvae per 45·cm branch<br />

tip. Application should start at shoot flare and continue no later than the date when 30% <strong>of</strong><br />

the spruee budworm have developed to the fifth instar.<br />

Deposit specifications - agar plates or Millipore® filters should be used to collect ground<br />

level deposits. aiming for a density <strong>of</strong> 25 droplets/cm l or more.<br />

In an attempt to reduce the inconsistencies <strong>of</strong> results in previous years, the 8.t.<br />

applications using these guidelines were co-ordinated in 1979 and 1980 (Morris 1980,<br />

1981). In general, the guidelines were followed unless extenuating circumstances were<br />

encountered. Clearly, B.I. is an effective alternative to the use <strong>of</strong> chemical pesticides for<br />

the protection <strong>of</strong> foliage. An analysis <strong>of</strong> the results <strong>of</strong> these applications (Morris 1980)<br />

showed that successful treatments had the following common characteristics:<br />

1. Pre-spray densities were usually less than 28 larvae per 45-cm branch tip<br />

2. Larval development at spray time was around fourth instar<br />

3. Bud flushing was 80-100% complete (except on red spruce)<br />

4. LV. applied per hectare were 20-40 x 10" in single or double applications<br />

5. Ground level droplet density was greater than 25/cm 1<br />

6. Spray time relative humidity was mostly higher than 65%<br />

7. Good weather followed spraying.<br />

These conditions are now generally accepted by researchers as desirable for 8.1.<br />

application in the forest.<br />

Smirn<strong>of</strong>f (1980c) advocates further conditions for reducing costs and increasing success;<br />

a) The required 8.1. dosage should be emitted in a formulation dispensible at<br />

2.5-3.0 l/ha; or (in the case or Thuricide® 32B with sorbitol in water) at 4.711ba;<br />

b) formulations should include Chevron® sticker at 1:1600; and chitinase at 10 000<br />

nephelometric units per hectare;<br />

c) spore viability in the concentrate drum should be checked;<br />

d) large capacity aircraft equipped with boom and nozzle are more cost-effective<br />

than small aircraft. Spray output must be calibrated and checked by counts <strong>of</strong> viable<br />

spores at ground level; the ground deposit should attain at least 60% <strong>of</strong> the emission<br />

rate;<br />

e) B.t. treatments should be carried out when most <strong>of</strong>the larvae are third instar, and<br />

buds are flushing (Smirn<strong>of</strong>f 1980c. Auger el 01. 1981).<br />

Improvements in commercial formulations and application technology during the past<br />

decade have brought B.I. to a point where it is now considered a clear alternative to<br />

chemical insecticides for use against spruce budworm. especially in environmentally<br />

sensitive areas. It is almost as effective for tree protection as the available chemical<br />

pesticides. Success or failure cannot yet be totally explained or predicted because<br />

relationships between foliage protection and dosage per volume emitted are not firmly<br />

established. Moreover, estimates <strong>of</strong> ground deposits have borne little practical relationship<br />

to deposits on foliage. The technical guidelines prepared in 1978 (see earlier)


246 W. A. Smirn<strong>of</strong>f and O. N. Morris<br />

Literature Cited<br />

continue to be valid, but require updating. B.t. spray operations are expensive<br />

compared with the use <strong>of</strong> chemical insecticides (2.5-5 times higher around 1980). As<br />

much <strong>of</strong> the cost is in material transport, more concentrated proprietary products should<br />

be devised to lower the shipping and application costs.<br />

Technical development <strong>of</strong> formulations, spray applications, and efficacy assessment<br />

should be continued: the search for more efficaceous formulations, droplet spectra, and<br />

timings should be continued; field dosage and volume/response relationships should be<br />

established, to provide the pest manager with costlbenefit options; and a quantitative<br />

method <strong>of</strong> measuring deposits at the feeding site is needed, to relate deposit rate <strong>of</strong> the<br />

active ingredient to its effectiveness.<br />

Auger. M.; Chabot. M.; Pelletier. M.; Bordeleau. C. (1981) Ellpertises entomologiques relices aux pulvl!risations al!riennes contre la<br />

tordeuse des bourgcons de I'!!pinette au QuI!bec en 1980. Qucbec; Ministl!re de I'Energie et des Ressources,<br />

107 pp.<br />

Cooksey, K.E. (1971) The protein erystaltollin <strong>of</strong> Bacillus thuringiensis: biochemistry and mode <strong>of</strong>action.ln: Burges. H.D.; Hussey. N.W.<br />

(Eds.) Microbial control <strong>of</strong> insects and mites. New York; Academic Press, pp. 247-274.<br />

Dorais, L.; Pelletier, M.; Smirn<strong>of</strong>f, W.A. (1980) Pulvl!risations acriennes de Bacillus thuringiensis Berliner realisees au Qucbec de 1971 a<br />

1979 contre In tordeuse des bourgeons de I'cpinette, Choristoneura fumiferana (Clem.). Rappon presente<br />

au VI' Congres International de l'Aviation Agricole a Turin. Italie, Septembre 1980.<br />

Fast. P.G. (1981) The crystal toxin <strong>of</strong> Bacillusthuringiensis.ln: Burges, H.D. (Ed.). Microbial control <strong>of</strong> pests and plant diseases 1970-1980.<br />

New York; Academic Press, pp. 223-248.<br />

Morris, O.N. (1973) Dosage·monality studies with commercial Bacillus thuringiensis sprayed in a modified Potler's Tower against some<br />

forest insects. Journal <strong>of</strong> Invertebrate Pathology 22,108-114.<br />

Morris, O.N. (1980) Repon <strong>of</strong> the 1979 CAN USA Cooperative Bacillus thuringiensis (B.t.) spray trials. Canadian Forestry Service<br />

Information Report FPM-X-40, 160 pp.<br />

Morris, O.N. (1981) Repon on the 1980 Cooperative Bacillus thuringiensis (B.t.) spray trials. Canadian Forestry Service Information Report<br />

FMP-X-48, 80 pp.<br />

Morris, O.N.; Moore. A. (1975) Studies on the protection <strong>of</strong> insect pathogens from sunlight protection. II. Preliminary field trials. Canadian<br />

Forestry Service Information Report CC-X-113.<br />

Morris, O.N.; Angus, T.A.;Smirn<strong>of</strong>f, W. (1975) In: Prebble, M.L. (Ed.). Aerial control <strong>of</strong>forest insects in <strong>Canada</strong>. Ottawa; Thorn Press, pp.<br />

129-133.<br />

Morris, O. N.; Hildebrand, M.J.; Moore, A. (1981) Comparative effectiveness <strong>of</strong> three commercial formulations <strong>of</strong> Bacillus thuringiensis for<br />

control <strong>of</strong> the spruce budworm, Choristoneura fumiferana (Oem.). Canadian Forestry Service<br />

Information Report FMP-X-47. 31 pp.<br />

Smirn<strong>of</strong>f, W.A. (1967) Innuence <strong>of</strong> temperature on the rate <strong>of</strong> development <strong>of</strong> sill varieties <strong>of</strong> Bacillus cereus group. Insect pathology and<br />

microbial control. Amsterdam; Nonh-Holland, pp. 125-130.<br />

Smim<strong>of</strong>f, W.A. (1973) Results <strong>of</strong> tests with Bacillus thuringiensis and chitinase on larvae <strong>of</strong> the spruce budworm. Journal <strong>of</strong> Invertebrate<br />

Pathology 21,116-118.<br />

Smirn<strong>of</strong>f, W.A. (1977) Confirmations ellperimentales du potential du complelle Bacillus thuringiensis et chitinase pour la rl!pression de la<br />

tordeuse des bourgeons de I'cpinette Choristoneura fumiferana (Lepidoptera: Tortricidae). Canadian<br />

Entomologist 109.351-358.<br />

Smim<strong>of</strong>f, W.A. (1978) Impact des traitemcnts bacteriologiques et des traitcmcnts chimiqucs sur une I!pidemie de la tordeuse des bourgeons<br />

de I'cpinette. Resumc des communications, 46' congres ACFAS. p. 68.<br />

Smirn<strong>of</strong>f, W.A. (1979a) Results <strong>of</strong> spraying Bacillus thuringiensis two consecutive years over balsam fir stands damaged by spruce<br />

budworm. Canadian Journal <strong>of</strong> Forest Research 9,509-513.<br />

Smirn<strong>of</strong>f, W.A. (1979b) Result <strong>of</strong> a 3-phase research program related to the biological control <strong>of</strong> the spruce budworm in 1979. LFRC, Repon<br />

to the Candian Forest Pest Control Forum, Nov. 27-28, 1979, Ottawa, Ontario.<br />

Smirn<strong>of</strong>f. W.A. (1979c) Changes in metabolites in insects during starvation, infections or poisoning. Abstracts <strong>of</strong> papers, 9th International<br />

Congress <strong>of</strong> Plant Protection and 71st Annual Meeting <strong>of</strong> the American Phytopathological <strong>Society</strong>,<br />

Washington, D.C., No. 928.<br />

Smirn<strong>of</strong>f, W.A. (19800) Deposit asscssment <strong>of</strong> Bacillus thuringiensis formulations applied from an aircraft. CalUJdian Journal <strong>of</strong><br />

Microbiology 26,1364-1366.<br />

Smim<strong>of</strong>f, W.A. (1980b) Calibration tests with various Bacillus thuringiensis preparations. Canadian Forestry Service Bi-monthly Research<br />

Notes 36,30-31.<br />

Smirn<strong>of</strong>f, W.A. (198Oc) Technological requirements for a successful Bacillus thuringiensis application against spruce budworm. Canadian<br />

Forestry Service Bi-monthly Research Notes 36.29-30.<br />

Smirn<strong>of</strong>f, W.A. (19810) Developpement d'une preparation compacte et economique de Bacillus thuringiensis pour la rl!pression de la<br />

tordeuse des bourgeons de I'cpinetle Choristoneura fumiferana. Annales de I'ACFAS 48.203.


248 J. C. Cunningham and G. M. Howse<br />

Background<br />

Field Trials<br />

B. Viruses: Application and Assessment<br />

J.C. CUNNINGHAM and G.M. HOWSE<br />

Five different types <strong>of</strong> viruses have been found to infect larvae <strong>of</strong> the spruce budworm,<br />

Chorisloneura fumiferana (Clem.). Four <strong>of</strong> them share one feature in common, that is,<br />

virus particles are contained within proteinaceous inclusion bodies. These infectious<br />

particles are released when the inclusion bodies are ingested by larvae and the inclusion<br />

body protein dissolves in their alkaline gut juices. The four viruses comprise a nuclear<br />

polyhedrosis virus (NPV), a granulosis virus (GV), a cytoplasmic polyhedrosis virus<br />

(CPV), and an entomopoxvirus (EPV). Between 1971 and 1980, all four viruses have<br />

been field tested either alone or in various combinations. The fifth virus, which has been<br />

found to infect spruce budworm, is cricket paralysis virus. It has no inclusion body, has<br />

a very wide host range, has little potential as a biological control agent, and will not be<br />

discussed further. Collapse <strong>of</strong> a spruce budworm population due to a naturally occurring<br />

virus epizootic has never been observed. Both NPV and CPV are occasionally detected<br />

in eastern spruce budworm populations in Ontario. but generally less than 1 % <strong>of</strong> larvae<br />

are infected.<br />

The first field trials <strong>of</strong> spruce budworm viruses were conducted in 1959 and 1960 by<br />

applying NPV and GV on single small trees (Stairs & Bird 1962). In 1969, different<br />

dosages <strong>of</strong> NPV were applied on single trees at 2- to 3-day intervals throughout the larval<br />

period (Bird & McPhee 1970). In 1970, an EPV was found in 2-year cycle spruce<br />

budworm, C. biennis Freeman (Bird el al. 1971). In the laboratory very low dosages <strong>of</strong><br />

this EPV (in comparison to other viruses) caused high larval mortality and, although it<br />

was slow acting, there was high hope that it would prove to be an effective biological<br />

control agent. By 1970, the use <strong>of</strong> artificial diet had greatly improved methods <strong>of</strong> rearing<br />

spruce bud worm (McMorran 1965), and sufficient larvae could be reared in order to<br />

produce viruses for aerial spray trials (Grisdale 1970).<br />

The first aerial spray trial was conducted in 1971 and trials have been carried out each<br />

season since then. Sixty plots were treated with viruses or combinations <strong>of</strong> chemical and<br />

viral insecticides in Ontario between 1971 and 1980. Such factors as dosage, timing <strong>of</strong><br />

application, tank mixtures, spray equipment, impact on different species <strong>of</strong> host trees,<br />

and persistence <strong>of</strong> viruses from one year to the next have been evaluated. The following<br />

viruses and combinations <strong>of</strong> viruses were tested during this period: NPV, EPV, GV,<br />

EPV plus NPV plus CPV, and NPV plus CPV. In addition, NPV and EPV plus NPV have<br />

been applied following chemical insecticide sprays.<br />

Assessment <strong>of</strong> the impact <strong>of</strong> treatments<br />

The same methods used to evaluate chemical insecticide treatments in Ontario were<br />

used to assess virus spray trials. In addition to carrying out larval population studies<br />

and defoliation estimates, attempts were made to determine the levels <strong>of</strong> infection in<br />

the spruce bud worm population, to assess the effect on successful pupal emergence and,<br />

on some plots. to continue these studies for one or more years after the initial application.


250 J. C. Cunningham and G. M. Howse<br />

Estimating current year's defoliation<br />

The percentage <strong>of</strong> the current year's defoliation was obtained by detailed examination <strong>of</strong><br />

the 46-cm branch tips collected for the post-spray sample from the treated and check<br />

plots. These samples were collected after all surviving larvae had pupated and feeding<br />

had ceased. In the tables, the percentage <strong>of</strong> foliage saved was calculated using Abbott's<br />

formula.<br />

Follow-up studies<br />

Many plots were re-examined the following year and some plots were studied for several<br />

years because carry-over <strong>of</strong> viruses from one year to the next and the impact on the spruce<br />

budworm population in the years following the year <strong>of</strong> treatment are considered important<br />

factors in evaluating these biological control agents. The same observations were recorded<br />

as in the year <strong>of</strong> treatment.<br />

Tank mixes for virus applications<br />

Several tank mixes have been used for spruce budworm virus applications and are described in<br />

the tables. The viral product in all tests was a finely ground lyophilized preparation <strong>of</strong><br />

virus-infected larvae. A few plots were treated with this product in water alone, but<br />

without a tracer dye no deposit assessment could be made. In many <strong>of</strong> the tests IMC 90-<br />

001 (International Minerals Corp.) UV protectant was added. It is a water-soluble wood<br />

byproduct that has a dark colour; it can also be used as a tracer dye. This material was<br />

renamed Shade® (Sandoz Inc.), but its production was discontinued in 1979.<br />

Molasses has been used in most <strong>of</strong> the tank mixes. It is an anti-evaporant that enhances<br />

deposit <strong>of</strong> aqueous sprays, may give some UV protection, and may be a feeding attractant<br />

to spruce budworm larvae. The results from single-tree ground spray tests indicated that<br />

molasses in the mix had a marked beneficial effect. A commercial molasses preparation,<br />

Sandoz adjuvant V® (Sandoz Inc.), was used in 1975.<br />

Many wetting and sticking agents have been tested in combination with Shade®, with<br />

or without molasses. These agents include Bi<strong>of</strong>ilm® (Colloidal Products Corp.), Chevron®<br />

spray sticker (Chevron Chemical Co.), Triton® X-lOO and Triton® B-1956 (Rohm and<br />

Haas Co. <strong>of</strong> <strong>Canada</strong> Ltd.). All passed a laboratory test to determine their compatibility<br />

with virus at a concentration <strong>of</strong> 1 %.<br />

Two other preparations tested were:Sandoz San 285® wettable powder, which is a<br />

feeding attractant for the virus <strong>of</strong> cotton bollworm, Heliothis zea (Boddie), incorporated<br />

in a polymer-bound carbon formulation prepared by the Southwest Research Institute,<br />

San Antonio, Texas; and an emulsifiable oil, Sunspray® 11E (Sunoco Inc.). The process<br />

used to produce the polymer-bound carbon formulation inactivated most <strong>of</strong> the virus<br />

and the feeding attractant did not enhance virus infection levels in spruce bud worm<br />

(Cunningham et al. 1978). The emulsifiable oil used at 66% had no adverse effect on the<br />

virus and its use would be beneficial if sprays were applied when the humidity was low.<br />

Entomopoxvirus plus nuclear polyhedrosis virus plus cytoplasmic polyhedrosis virus<br />

In 1971 a helicopter was used to treat six 1.07-ha plots with EPV (Table 63). When<br />

smears <strong>of</strong> larvae from treated plots were examined microscopically it was discovered<br />

that the EPV preparation was contaminated with both NPV and CPV. All three viruses<br />

gave consistently higher levels <strong>of</strong> infection in larvae on white spruce, Picea glauca<br />

(Moench) Voss, hosts than on balsam fir, Abies balsamea (L.) Mill., hosts (Bird et al.<br />

1972, Howse et al. 1973).


Table 63<br />

Table 64<br />

B. Viruses: Application Clnd assessment 251<br />

Application <strong>of</strong> EPV contaminated with NPV and CPV on six 1.07-ha plots using a helicopter<br />

fitted with boom and nozzle equipment and calibrated to deliver 28.2 Ilha*<br />

Population<br />

Pre·dominant Deposit Highest virus infection (01.,) reduction due Current<br />

Dosage. instar at card, to treatment year's foliage<br />

PIB/ha time <strong>of</strong> droplets! EPV NPV CPV (%} saved ('Yo}<br />

Plot (x 10") application cm 2 bF·· wS bF wS bF wS bF wS bF wS<br />

1 750 2 10 9 38 10 0 7 0 40 R 15<br />

2 75 2 9 3 16 0 12 1 4 0 61 3 0<br />

3 7.5 2 9 1 12 0 2 II II 0 25 0 0<br />

4 750 3&4 29 6 5 4 7 1 4 30 79 1 0<br />

5 75 3&4 31 2 2 0 1 II II 48 57 0 0<br />

6 7.5 3&4 NO··· 4 6 0 0 0 0 II 59 1 0<br />

* Aqueous tank mix contained 2.5% IMC 90-001 sunlight protectant.<br />

** bF = balsam fir, Abies balsamea (L.) Mill.; wS = white spruce, Picea glauca<br />

(Moench) Voss.<br />

••• ND = Not determined .<br />

Follow-up studies were undertaken between 1972 and 1975 on the two plots treated<br />

with the highest dosage (Table 64). It can be seen that, although there was little foliage<br />

protection in 1971 (Table 63), there was significant foliage protection in the subsequent 3<br />

years. Also, the NPV, which was a minor contaminant in the EPV preparation, was<br />

considerably more prevalent in the following years than the EPV (Cunningham el al.<br />

1975a).<br />

Follow-up studies on two I?lots treated with EPV contaminated with NPV and CPV at a<br />

dosage <strong>of</strong> 750 x to" inclUSIon bodieslha in 1971·<br />

Population Current<br />

Highest virus infection (%) reduction due "ear's<br />

to virus oliage<br />

EPV NPV CPV {%l saved {%}<br />

Plot Year bF** wS bF wS bF wS bF wS bF wS<br />

1972 0 5 7 19 2 5 64 82 72 59<br />

1973 2 5 3 13 0 0 3 60 27 22<br />

1974 1 2 1 4 0 0 0 5 29 49<br />

1975 1 6 5 16 1 1 0 78 22 2<br />

4 1972 0 0 17 20 to 4 16 68 44 49<br />

1973 0 1 4 16 2 0 15 83 26 22<br />

1974 0 1 3 11 1 0 0 48 25 40<br />

1975 0 1 17 11 0 0 47 26 1 1<br />

* See Table 63 for details and results in 1971<br />

.* bF = balsam fir, Abies balsamea (L.) Mill.; wS<br />

(Moench) Voss.<br />

Entomopoxvirus<br />

white spruce, Picea glauca<br />

In 1972, pure EPV was tested and three plots with a total area <strong>of</strong>512 ha were treated. The<br />

timing <strong>of</strong> the application was on second- and third-instar larvae, but an unseasonably<br />

late snow fall destroyed most <strong>of</strong> the spruce budworm larvae and killed the new growth<br />

on the trees. Population studies and defoliation estimates were abandoned and only


Table 67<br />

B. Viruses: Application and assessment 253<br />

levels <strong>of</strong> CPV as well as <strong>of</strong> NPV infection were recorded (Table 66) (Bird el al. 1972,<br />

Howse el al. 1973). Follow-up studies were undertaken on these plots between 1972 and<br />

1977 (Table 67). The incidence <strong>of</strong> NPV was higher than CPV during this period. Although<br />

not spectacular, some saving <strong>of</strong> foliage was recorded between 1972 and 1975 (Cunningham<br />

el al. 1975a).<br />

Follow-up studies on two white spruce, Picea glauca (Moench) Voss, plantations treated<br />

with NPV plus CPV (400:1) in 1971-<br />

Highest virus<br />

infection %<br />

Population<br />

reduction<br />

due to virus Current year's<br />

Plot Year NPV CPV carry-over ('Yo) foliage saved ('Yo)<br />

G 1972 28 8 20 0<br />

1973 17 5 82 22<br />

1974 6 I 56 27<br />

1975 21 0 74 0<br />

1976 8 0 NO-- NO<br />

1977 3 0 NO NO<br />

H 1972<br />

1973<br />

1974<br />

1975<br />

1976<br />

1977<br />

24<br />

14<br />

6<br />

9<br />

0<br />

2<br />

3<br />

0<br />

0<br />

0<br />

0<br />

0<br />

65<br />

68<br />

62<br />

48<br />

NO<br />

NO<br />

63<br />

20<br />

44<br />

0<br />

ND<br />

NO<br />

- See Table 66 for details and results in 1971.<br />

-. ND = not determined.<br />

Rigorous quality control facilitated production <strong>of</strong> pure NPV between 1972 and 1978, but<br />

in 1979 a CPV contaminant was again found and aUempts to eliminate it were unsuccessful<br />

both that year and in 1980. In 1979, the ratio <strong>of</strong> NPV:CPV was 178:1, and in 1980 the<br />

same preparation and one with an NPV:CPV ratio <strong>of</strong> 300: I were applied (W.J. Kaupp<br />

1980, personal communication) (Table 66). In 1979, the viruses were applied either as<br />

second-instar larvae emerged from hibemacula or on later instars following budflush.<br />

Following the early application, high levels <strong>of</strong> virus infection and acceptable foliage<br />

protection were recorded. On the remaining plots reasonable levels <strong>of</strong> virus infection<br />

and population reduction but little or no foliage protection were recorded. In 1980, an<br />

attempt was made to repeat the early applications made in 1979. However, when the<br />

spray was applied about 90% <strong>of</strong> the larvae had mined into needles and buds. Infection<br />

levels were lower than in 1979, although moderate levels <strong>of</strong> both popUlation reduction<br />

and foliage protection were recorded.<br />

Nuclear polyhedrosis virus<br />

Tests were conducted with pure NPV in 1972 (Cunningham & McPhee 1973), in 1973<br />

(Cunningham el al. 1974) in 1974 (Cunningham el al. 1975b), in 1975 (Cunningham etal.<br />

1975c), in 1976 (Kaupp el al. 1978), in 1977 (Cunningham el al. 1978), and in 1978 (Cunningham<br />

el af. 1979). During that 6-year period, 33 plots with a total area <strong>of</strong> 1741 ha were treated<br />

(Table 68). In 1972 the trials were a failure because <strong>of</strong> the same late snow fall that<br />

affected the EPV trials that year. In subsequent years, tests were made with lower<br />

dosages than those used in the original 1971 test, a variety <strong>of</strong> tank mixes containing


258 J. C. Cunningham and G. M. Howse<br />

Recommendations<br />

use will probably be restricted to high value stands and environmentally sensitive areas. If it<br />

can be demonstrated conclusively that treatment with a virus gives foliage protection for<br />

several years following the year <strong>of</strong> application, then the high initial cost <strong>of</strong> the treatment<br />

becomes more acceptable.<br />

The application <strong>of</strong> contaminated virus preparations in 1971 was accidental. The<br />

presence <strong>of</strong> a CPV contaminant in the NPV used in 1979 and 1980 was known. but the<br />

material was still applied on the basis <strong>of</strong> the 1971 results. The minute amounts <strong>of</strong> CPV<br />

gave dramatic infection levels in the spruce budworm population that were <strong>of</strong>ten higher<br />

than the levels <strong>of</strong> NPV. This was surprising because the ratios <strong>of</strong> NPV:CPV were 400: 1.<br />

300: 1. and 178: 1. It appears that there is a synergistic effect between NPV and CPV. but<br />

this would require detailed laboratory investigations to prove conclusively.<br />

Foliage protection in the year <strong>of</strong> application was demonstrated in 1979 following treatment<br />

with NPV plus CPV on second-instar larvae as they emerged from hibernacula. Treatment<br />

<strong>of</strong> second-instar larvae has many benefits. although timing is extremely critical and<br />

it would only be practical to treat small. carefully monitored areas. Second-instar larvae<br />

are much more susceptible to infection with NPV than later instars and the LD50 in terms<br />

<strong>of</strong> numbers <strong>of</strong> PIB has been calculated for different instars as follows: second. 25; third.<br />

204; fourth. 462; fifth, 2514; and sixth, 3 643 (F.T. Bird 1978, unpublished data). At the<br />

time <strong>of</strong> emergence from hibernacula. leaves on the hardwood overstorey. frequently<br />

found in spruce-fir stands, have not flushed and a good aerial spray deposit can be<br />

obtained on the target species.<br />

To date, the tank mix considered most suitable for spruce budworm virus applications<br />

was an aqueous suspension containing 25% molasses, 6% Shade® and 1 % Chevron®<br />

sticker. It has been widely used with the NPVs <strong>of</strong> both Douglas-fir tussock moth. Orgyia<br />

pseudotsugata (McDunnough), and gypsy moth, Lyman/ria dispar (L.). Since Shade® is<br />

no longer commercially available, it will be necessary to find a new screening agent<br />

against ultra-violet light or a totally new formulation.<br />

Safety tests on spruce budworm NPV conducted at Ontario Veterinary College<br />

showed that this virus presents no hazard to birds and mammals (Valli et al. 1976). Fish<br />

tests, although indicating no hazard. were inconclusive and should be repeated (Savan<br />

et al. 1979). There are only limited safety testing data available for EPV (Buckner &<br />

Cunningham 1972). and no CPV and GV safety testing data are available. These would<br />

be required before any <strong>of</strong> these three .. iruses could be registered by Canadian authorities.<br />

The use <strong>of</strong> any <strong>of</strong> the four viruses with inclusion bodies that infect spruce budworm is still at<br />

the experimental stage. In the near future only treatment <strong>of</strong> small areas can be considered.<br />

Applications <strong>of</strong> NPV or NPV plus CPV on second-instar larvae as they emerge from<br />

hibemacula should be retested in an effort to find if. under closely monitored conditions.<br />

this is a feasible and effective strategy.<br />

An improved tank mix that will keep the virus in a viable state on the foliage for more than<br />

a week would greatly increase the effectiveness <strong>of</strong> treatments and possibly allow lower<br />

dosages. Research on formulations is being conducted by both commercial and government<br />

agencies in the United States and formulations developed for other pathogens. such as<br />

bacteria and fungi. will probably have the characteristics suitable for use with viruses.<br />

The search for new viruses or strains <strong>of</strong> viruses that infect spruce budworm should be<br />

continued. There are several isolates <strong>of</strong> NPV • GV, and EPV from different Choristoneura<br />

species, but none has been found that is more virulent than the strains used in field trials.<br />

The search for larger. easily reared host larva for spruce budworm virus production has also<br />

proved unsuccessful. but should be continued.<br />

A long-term goal in the use <strong>of</strong> viruses for the regulation <strong>of</strong> spruce budworm population is<br />

to study the nucleic acid <strong>of</strong> spruce budworm NPV with a view to constructing a genetic


Literature Cited<br />

B. Viruscs: Applicmion and assessmcnt 259<br />

map. Once this is accomplished, techniques <strong>of</strong> genetic manipulation can be used in an<br />

attempt either to increase the virulence <strong>of</strong> spruce budwonn NPV or to modify some highly<br />

virulent, easily produced insect virus in such a way as to render it capable <strong>of</strong> infecting<br />

spruce budwonn.<br />

Abbott. W.S. (1925) A method <strong>of</strong> computing the effectiveness <strong>of</strong> an insecticide. Journal <strong>of</strong> Economic Entomology 18.265-267.<br />

Arif. B.M.; Sohi. S.S.; Krywienczyk. J. (1976) Replication <strong>of</strong> alkali-released NPV in a cell line <strong>of</strong> ChorislOneurafumiferana. Proceedings <strong>of</strong><br />

the 1st International Colloquium on Im'mebrate Pathology Kingston, Ontario, pp. 108-110.<br />

Bird. F.T.; McPhee. J. R. (1970) Susceptibility <strong>of</strong> spruce budworm to pure nuclear polyhedrosis virus (NPV) sprays. Canadian Department<br />

<strong>of</strong> Fisheries and Forestry Bi·monthly Research Notes 26.35.<br />

Bird, F.T.; Sanders. C.J.; Burke. J.M. (1971) A newly discovered virus disease <strong>of</strong>the spruce budworm. Choristoneura biennis (Lepidoptera:<br />

Tortricidae). Journal <strong>of</strong> Invertebrate Pathology 18.159-161.<br />

Bird. F.T.; Cunningham. J.e.; Howse. G.M. (1972) Possible use <strong>of</strong> viruses in the control <strong>of</strong> spruce budworm. Proceedings <strong>of</strong> the<br />

<strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> Ontario 103,69-75.<br />

Buckner. C.H.; Cunningham. J.C. (1972) The effect <strong>of</strong> the poxvirus <strong>of</strong> the spruce budworm. Choristoneura fumiferana (Lepidoptera:<br />

Tortricidae). on mammals and birds. Canadian Entomologist 104.1333-1342.<br />

Cunningham. J.e.; McPhee. J.R. (1973) Aerial application <strong>of</strong> entomopoxvirus and nuclear polyhcdrosis virus against spruce budworm at<br />

Chapleau. Ontario, 1972. Canadian Forestry Service, Sault Ste. Marie, Information Report IP-X-3. 27 pp.<br />

Cunningham, J.C.; McPhee, J.R.; Howse. G.M.; Harnden, A.A. (1974) Aerial application <strong>of</strong> a nuclear polyhedrosis virus against spruce<br />

budworm near Massey and Aubrey Falls, Ontario in 1973 and a survey <strong>of</strong> the impact <strong>of</strong> the virus in 1974.<br />

Canadian Forestry Service, Sault Ste. Marie, Information Report IP-X-4, 45 pp.<br />

Cunningham, I.C.; Howse, G.M.; Kaupp, W.J.; McPhee, J.R.; Hamden, A.A.; White, M.B.E. (1975a) The persistence <strong>of</strong> nuclear<br />

polyhedrosis virus in populations <strong>of</strong> spruce budworm. Canadian Forestry Service, Sault Ste. Marie,<br />

Information Report IP-X-lO, 32 pp.<br />

Cunningham. J.C.; Kaupp. W.J.; McPhee, I.R.; Howse, G.M.; Harnden, A.A. (1975b) Aerial application <strong>of</strong> nuclear polyhedrosis virus<br />

against spruce budworm on Manitoulin Island, Ontario in 1974 and a survey <strong>of</strong> the impact <strong>of</strong> the virus in<br />

1975. Canadian Forestry Service, Sault Ste. Marie, Information Report IP-X·8, 30 pp.<br />

Cunningham. I.C.; Kaupp. W.I.; McPhee, J.R.; Howse, G.M.; Harnden, A.A. (1975c) Aerial application <strong>of</strong> nuclear polyhcdrosis virus on<br />

spruce budworm on Manitoulin Island, Ontario in 1975. Canadian Forestry Service, Sault Ste. Marie,<br />

Information Report IP-X-9, 35 pp.<br />

Cunningham. J.c.; Kaupp. W.J.; Howse. G .M. ; McPhee. J. R.; de Groot. P. (1978) Aerial application <strong>of</strong> spruce budworm baculovirus: tests<br />

<strong>of</strong> virus strains. dosages and formulations in 1977. Canadian Forestry Service. Sault Ste. Marie.<br />

Information Report FPM-X-3. 37 pp.<br />

Cunningham. I.C.; Howsc. G.M.; McPhee. J.R.; de Groot. P.; White. M.B.E. (1979) Aerial application <strong>of</strong> spruce budworm baculovirus:<br />

replicated tests with an aqueous formulation and a trial using an oil formulation in 1978. Canadian<br />

Forestry Service, Sault Ste. Marie, Information Report FPM-X-21. 19 pp.<br />

Grisdale. D. (1970) An improved method for rearing large numbers <strong>of</strong> spruce budworm. Chonstoneura fumiferana (Lepidoptera: Tonricidae).<br />

Canadian Entomologist 102.1111-1117.<br />

Howse. G.M.; Sanders. C.J.; Harnden. A.A.; Cunningham, J.e.; Bird, F.T.; McPhee. J.R. (1973) Aerial application <strong>of</strong> viruses against<br />

spruce budworm. 1971. Pan A. Impact in the year <strong>of</strong> application (1971). Pan B. Impact in the year<br />

following application (1972). Canadian Forest,)' Service, Sault Ste. Marie. Information Report O-X·I89.<br />

62pp.<br />

Kaupp. W.J. Cunningham,; J.C.; Howse. G.M.; McPhee. J.R.; de Groot. P. (1978) Aerial application <strong>of</strong>spruce budworm baculovirus: tests<br />

on sccond instar larvae in 1976. Canadian Forestry Service. Sault Ste. Marie. Information Report FPM­<br />

X-2. 20 pp.<br />

McMorran. A. (1965) A synthetic diet for the spruce budworm. Chonstoneura fumiferana (Clem.) (Lepidoptera: Tonricidae). Canadian<br />

Entomologist 97,58-62.<br />

Morris. O.N.; Armstrong. J.A.; Hildebrand. M.J.; Howse. G.M.; Cunningham, J.C.; McPhee. J.R. (1972) Aerial applications or virusinsccticide<br />

combinations against spruce budworm Chonstontura fumifl'rana (Clem.) (Tonricidae:<br />

Lepidoptera) at Rankin. Ontario, 1972. Canadian Forestry Servicl', Ollawa, Information Report CC-X-<br />

37,65 pp.<br />

Morris. O.N.; Armstrong, J.A.; Howse. G.M.; Cunningham, J.e. (1974) A 2-year study <strong>of</strong> virus-chemical insecticide combination in the<br />

integrated control or the spruce budworrn. Choristoneura fumiferana (Tortrieidae: Lepidoptera).<br />

Canadian Entomologist 106.813-824.<br />

Savan. M.; Budd. J.; Reno, P.W.; Darley, S. (1979) A study <strong>of</strong> two species <strong>of</strong> fish inoculated with spruce budworm nuclear<br />

polyhedrosis virus. Journal <strong>of</strong> Wildlife Disl'ases 15,331-334.<br />

Stairs, G.R.; Bird. F.T. (1962) Dissemination <strong>of</strong> viruses against the spruce budworm. Choristoneura fumiferana (Clemens). Canadian<br />

Entomologi.ft 94,966-969.<br />

Valli, V.E.; Cunningham, J.C.; Arir, B.M. (1976) Tests demonstrating the safety or a baculovirus <strong>of</strong> the spruce budworm to mammals and<br />

birds. Proceedings <strong>of</strong> the 1st International Colloquium on Invertebrate Patllology, Kingston, Ontario, pp.<br />

445-446.<br />

Wilson, G.G. (1981) Nosema fumiferanae, a natural pathogen <strong>of</strong> a forest pest: potential ror pest management. In: Burges. H.D. (Ed.)<br />

Microbial control <strong>of</strong> pests and plant diseases 1970-1980. London; Academic Press. pp. 595-601.


C. AppliclItion <strong>of</strong> microsporidia and fungi. 261<br />

The natural levels <strong>of</strong> microsporidian infection in spruce budworm populations tend to<br />

increase with the age <strong>of</strong> the infestation. Examination <strong>of</strong> spruce budworm larvae collected in<br />

the Uxbridge forest <strong>of</strong> southern Ontario over a 6-year period indicated that the infection<br />

rate from N.fllmiferanae increased from 36% in 1973 to 69% in 1978 (Wilson 1977b). The<br />

Uxbridge forest was in fact an old infestation; severe defoliation was recorded in 1952<br />

and has persisted at fluctuating levels ever since.<br />

Fungi<br />

Steady progress was made during the 1970s in documenting the effects <strong>of</strong> fungi in natural<br />

epizootics and developing various species for use in biological control, with the result<br />

that Hirslllelia thompsoni Fisher, the first fungus to be registered for mite control in<br />

North America, is now available under the trade name Mycar® (Abbott Laboratories)<br />

and is also used against citrus rust mites.<br />

The occurrence <strong>of</strong> entomophthoraceous fungi on spruce budworm has been documented<br />

throughout much <strong>of</strong> the insect's range in the last 10 years. Outbreaks <strong>of</strong> fungal disease<br />

occurred in 1974 and 1979 in northern Ontario; in the former year the outbreak was <strong>of</strong><br />

Zoophlhora radicans (Brefeld) A. Bakto (= Entomophthora sphaerosperma Frasenius)<br />

(Harvey & Burke 1974), in the latter it was <strong>of</strong> the same fungus together with E. egressa<br />

MacLeod (Tyrrell, unpublished). In Newfoundland, the same two fungi were first<br />

recorded from the spruce budworm in 1972. The disease was widespread in western<br />

Newfoundland in 1973, and subsequently spread to the whole <strong>of</strong> the Island (Otvos &<br />

Moody 1978). Mortality varied from 10-40%, depending on weather conditions, and<br />

occasionally reached much higher values in localized areas. Both fungal species, together<br />

with an unidentified Conidiobo/w species, were responsible for spuce budworm mortality<br />

in northern Maine, where disease prevalence varied between 4 and 7% during 1975-77<br />

(Vandenberg & Soper 1978).<br />

In northern Ontario, mortality was higher in white spruce than in balsam fir in 1974,<br />

when fungal incidence was high (Harvey & Burke 1974), but no difference was observed<br />

between tree species in 1980, when fungal incidence was much lower (TyrreU, unpublished). In<br />

Maine, fungal prevalence was found to be both density-dependent and significantly<br />

higher in the lower crown area <strong>of</strong> the tree, and appeared to be enhanced by periods <strong>of</strong><br />

cool, wet weather (Vandenberg & Soper 1978). The link between moist conditions and<br />

• increased incidence <strong>of</strong> fungus has also been observed in Newfoundland (Otvos &<br />

Moody 1978).<br />

Fungal infection occurs in the later stages <strong>of</strong> larval development. Incidence <strong>of</strong> Z.<br />

radicans peaked between the fifth and sixth instars, E. egressa between the sixth instar<br />

and pre-pupa, and Conidiobolus sp. between the pre-pupal and pupal stages, respectively,<br />

in Maine (Vandenberg & Soper 1978); while in northern Ontario peak fungal infection<br />

coincided with the peak development <strong>of</strong> sixth-instar larvae (Tyrrell unpublished).<br />

Earlier instars however, are not resistant to the disease, and can readily be infected with<br />

Z. radicans under laboratory conditions (Kenneth 1978, Perry unpublished).<br />

Both Z. radicans and E. egressa can be isolated and grown in pure culture, and a<br />

number <strong>of</strong> different isolates <strong>of</strong> both species are now available. Z. radicans grows readily<br />

on a variety <strong>of</strong> solid and liquid media, and the mycelial stage can also be grown in small<br />

fermentors. E. egressa grows slowly on complex media such as coagulated egg yolk, but<br />

can be grown as a protoplast in osmotically stabilized media.<br />

Z. radicans can sporulate on artificial media to give both conidia and resting spores.<br />

The effects <strong>of</strong> environmental conditions, including temperature, photophase, and nutrient,<br />

on conidial germination have been determined (van Roermund, personal communication).<br />

Maturation <strong>of</strong> laboratory-produced resting spores under controlled conditions has been<br />

shown to be a key factor in potentiating their subsequent germination (Perry & Tyrrell<br />

in press), and the effects <strong>of</strong> environmental conditions on their germination are currently


264 G. G. Wilson, D. Tyrrell and T. J. Ennis<br />

Recommendations<br />

schubergi spores resulted in infection rates <strong>of</strong> 64.8% for white spruce and 96.3% for<br />

balsam fir 19 days after spraying. There was no infection in larvae from the check area<br />

and no carry-over <strong>of</strong> the microsporidium to the following year. All samples showed<br />

higher levels <strong>of</strong> infection in spruce budworm taken from balsam fir.<br />

Based on laboratory results there is a relationship between levels <strong>of</strong> infection and<br />

mortality. As an example, spore concentrations <strong>of</strong> 10" spores/ml fed to second-instar<br />

larvae resulted in 100% mortality in up to 24 days, with 50% mortality occurring about 14<br />

days after infection (balsam fir buds were dipped in the spore suspension). As a<br />

comparison, spore concentration <strong>of</strong> 100/ml resulted in 61 % mortality over a period <strong>of</strong><br />

6-28 days. For fifth-instar larvae Hr spores/ml gave 75% mortality over 8-33 days,<br />

and 10" sporeslml resulted in only 5.6% mortality, which was similar to the controls.<br />

Fungi<br />

E. egressa does not sporulate well on artificial media and conidia are normally obtained<br />

from infected insect cadavers. These are easily produced by injection <strong>of</strong> insect larvae<br />

with protoplasts, which have been shown to be part <strong>of</strong> the natural life cycle <strong>of</strong> this fungus<br />

(Tyrrell 1977). Transmission <strong>of</strong> E. egressa from laboratory-infected to natural populations<br />

<strong>of</strong> spruce budworm under field conditions using conidia produced in this manner has<br />

been demonstrated (Lim el al. unpublished).<br />

Genetic manipulation<br />

Topical application <strong>of</strong> thiotepa or 250-300 Sv <strong>of</strong> Co'" irradiation induced complete<br />

sterility in male spruce budworm (Retnakaran 1970, 1971). Males sterilized with thiotepa<br />

remained fully competitive and could suppress fertility <strong>of</strong> caged laboratory populations<br />

by up to 90% (Ennis unpublished data). However, release <strong>of</strong> chemosterilized males into<br />

caged field populations produced no detectable reduction in popUlation, mainly because<br />

<strong>of</strong> adverse effects on adult male mating behaviour. High level irradiation <strong>of</strong> male pupae<br />

or adults reduces fitness and mating competitiveness, militating against their effective<br />

use. However, at 20-50 Sv, induced chromosome aberrations have a heritable effect on<br />

fertility <strong>of</strong> subsequent generations, with depression <strong>of</strong> fertility persisting for up to three<br />

generations in the laboratory (Ennis 19790). The potential <strong>of</strong> inherited semi-sterility has<br />

not yet been tested in field populations. Dispersal and behaviour <strong>of</strong> irradiated males<br />

have also been studied using pheromone re-trapping techniques (Ennis 1979b).<br />

Few microsporidian parasites have been investigated as possible control agents <strong>of</strong> pest<br />

insects. If we take the amount <strong>of</strong> research on Bacillus Ihuringiensis Berliner (B.I.)<br />

before it was accepted as an insecticide as being representative in the development <strong>of</strong> a<br />

biological control agent, then it is obvious that we have barely begun the task with<br />

microsporidia. However, there are indications that many microsporidia do have good<br />

potential as biological control agents that, given appropriate research and development,<br />

may one day be realized. There is still much research needed on the development <strong>of</strong><br />

microsporidia in a control plan for the spruce budworm. Thorough studies are required<br />

on the host-parasite relationships and the incidence and effects <strong>of</strong> the microsporidia in<br />

their natural setting. Although these parasites can be introduced into a population <strong>of</strong> the<br />

spruce budworm, their effect on insect numbers has not been quantified. The use <strong>of</strong><br />

these parasites in integrated control should be investigated, not only in conjunction with<br />

chemicals, but also with other spruce budworm pathogens.


Literature Cited<br />

C. Application <strong>of</strong>microspnridia and fungi. 265<br />

The level <strong>of</strong> research effort on fungi for control <strong>of</strong> spruce budworm has been low<br />

compared to the effort on such biological agents as the viruses and B.t., but the results<br />

obtained to date suggest that fungi playa significant role in the regulation <strong>of</strong> spruce<br />

budworm populations, and the recent demonstration that infectious diseases can in<br />

theory drive population cycles (Anderson & May 1980) will provide further impetus for<br />

studies designed to delineate the role <strong>of</strong> fungi in the spruce budworm ecosystem. To this<br />

end, extensive epizootiological investigations on the biotic and abiotic factors affecting<br />

the fungus-budworm interaction should be carried out. These, coupled with continuing<br />

laboratory studies on the production <strong>of</strong> fungal material suitable for field dissemination.<br />

and on the selection and mode <strong>of</strong> action <strong>of</strong>the most virulent strains <strong>of</strong> fungi. will. in the<br />

long term. determine how and under what circumstances fungi may be used in the<br />

manipulation <strong>of</strong> spruce budworm populations.<br />

At present it is difficult to draw definite conclusions on the potential <strong>of</strong> microsporidia<br />

and fungi in the management <strong>of</strong> the spruce budworm. However. because <strong>of</strong> their<br />

potential in a control programme. any comprehensive model <strong>of</strong> spruce budworm population-dynamics<br />

must embrace a thorough understanding <strong>of</strong> these pathogens.<br />

Research into the genetic manipulation <strong>of</strong> the spruce budworm to date has supplied<br />

the basic level <strong>of</strong> information required for further testing: methods for mass sterilization.<br />

effects on fertility <strong>of</strong> caged population, and dispersal and detection <strong>of</strong> released insects.<br />

However, the large numbers <strong>of</strong> insects and extensive areas <strong>of</strong> forest affected by epidemics<br />

<strong>of</strong> this pest suggest caution in trying to develop genetic manipulation for control <strong>of</strong><br />

outbreak populations. The logistics <strong>of</strong> application, in terms <strong>of</strong> numbers and area. and the<br />

uncertainty <strong>of</strong> success indicate that its greatest potential lies in controlling endemic<br />

population levels, where the released insects' ability to detect and mate with females<br />

even when they are relatively rare provides one <strong>of</strong> the greatest advantages <strong>of</strong> the genetic<br />

approach. Future research should be aimed at such low level populations, and should be<br />

conducted with the objective <strong>of</strong> reducing reproduction potential to a level below that<br />

necessary to support a transition from endemic to epidemic levels.<br />

Anderson. R.M.; May. R.M. (1980) Infectious diseases and population cycles <strong>of</strong> forest insects. Science 210.658-661.<br />

Burges. H.D. (Ed.) (1981) Microbial control <strong>of</strong> pests and plant diseases, 1970-1980. London; Academic Press, 949 pp.<br />

Ennis. TJ. (1979a) Detection and isolation <strong>of</strong> induced chromosome aberrations in Lepidoptera. Experientia 35.1153-1154.<br />

Ennis, T.J. (1979b) A release·recapture experiment with normal and irradiated spruce budworrn males. Cana4ian Forestry Service Bi·<br />

monthly Research Notes 35.9-10.<br />

Harvey. G.T.; Burke. J.M. (1974) Monality <strong>of</strong> the spruce budworrn on white spruce caused by Entomophthora sphaerosperma. Canadian<br />

Forestry Service Bi·monthly Research Notes 30.23-24.<br />

Kenneth. R.G. (1978) Entomophthora sphoerosperma as an insect pathogen for spruce budworrn control. Abstracts <strong>of</strong> the Acadian<br />

<strong>Entomological</strong> <strong>Society</strong>'s 38th AnnWlI Meeting. pp. 35-36.<br />

Lachana:. LE. (1979) Genetic strategies affecting the swx:css and economy <strong>of</strong> the sterile insect release method. In: Hoy. M.A.; McKelvey. J.J.<br />

(Eds.) Genetics in relation to insect management. Rockefeller Foundation. USA.<br />

Otvos.I.S.; Moody. B.H. (1978) The spruce budworm in Newfoundland: history. status and control. Canadian Foresty Sen·ice. Newfoundland<br />

Forest Research Centre, Information Report N·X-150.<br />

Outram. J. (1969) Potential use <strong>of</strong> the sterility principle for spruce budworm control in eastern <strong>Canada</strong>. Canadian Department <strong>of</strong> Fisheries and<br />

Fort'stry' InteTllal Report M-45. Fredericton. N.B.<br />

Perry. D.; Tyrrell. D. (in prcss) Resting spore germination in ZoophtllOra radicans. Mycologia.<br />

Rctnakaran. A. (1970) Preliminary results <strong>of</strong> radiation induced sterility <strong>of</strong> the male spruce budworm. Canadian Department <strong>of</strong> Fisherit's and<br />

Forestry' Bi·monthly Research Notes 26.13-14.<br />

Retnakaran. A. (1971) Thiotepa as an effective agent for mass sterilizing the spruce budworrn. Choristoneura fumiferana (Lepidoptera:<br />

Tortricidae). Canadian Entomologist \03.1753-1756.<br />

Thomson. H.M. (1958) The effect <strong>of</strong> a microsporidian parasite on the development. reproduction and mortality <strong>of</strong> the spruce budworrn.<br />

Choristoneura fumiferana (Clem.). Canadian Journal <strong>of</strong> Zoology 36.499-51 J.<br />

Tyrrell. D. (1977) Occurrence <strong>of</strong> protoplasts in the natural life cycle <strong>of</strong> Entomophthora egressa. Experimental Mycology 1.259-263.<br />

Vandenberg. J.D.; Soper. R.S. (1978) Prevalence <strong>of</strong> entomophthorales mycoses in populations <strong>of</strong> the spruce budworm. Choristoneura<br />

fumiferana. Environmental Entomology 7.847-853.


266 G. G. Wilson. D. Tyrrrell and T. J. Ennis<br />

Vandenberg, J.D.; Soper, R.S. (1979) A bioassay technique for Entomophthora sphaerosperma on the spruce budworm, Choristoneura<br />

fumiferana. Journal <strong>of</strong> Invertebrate Pathology 33,148-154.<br />

Wilson, G.G. (l977a) The effects <strong>of</strong> feeding microsporidian (Nosema fumiferanae) spores to naturally infected spruce budworm (Choristoneura<br />

fumiferana). Canadian Journal <strong>of</strong> Zoology 55,249-250.<br />

Wilson. G.G. (l977b) OhselVations on the incidence rates <strong>of</strong> Nosema fiuniferOlUle(microsporidia) in a spruce budwonn. ChoristoTU!ura fumiferana.<br />

(Lepidoptem: Tortricidae) population. Proceedings <strong>of</strong> Iile <strong>Entomological</strong> Sociely <strong>of</strong> Ontario 108,144-145.<br />

Wilson, G.G. (1980) Persistence <strong>of</strong> microsporidia in populations <strong>of</strong>the spruce budworm and forest tent caterpillar. Canadian Forestry Service,<br />

Sault Ste. Marie, Information Report FPM-X-39.<br />

Wilson, G.G. (1981) The potential <strong>of</strong> Pleistophora schubergi in microbial control afforest insects. Canadian Forestry Service, Sault Ste. Marie,<br />

Information Report FPM-X-49.


Background<br />

Table 73<br />

D. Testing <strong>of</strong> Parasitoids<br />

I.W. VARTY<br />

D. Testing <strong>of</strong> parasitoids 267<br />

The two strategies for biological control <strong>of</strong> the spruce budwonn, CJwristoneura fwniferatUl<br />

(Clem.), by parasitoids are the introduction <strong>of</strong> exotic species and the manipulation <strong>of</strong><br />

native ones. Most entomologists agree that the best prospects for introduction are those<br />

species that parasitize Choristoneura spp. (or near relatives) in spruce (Picea spp.) and fir<br />

(Abies spp.) forests in similar climatic zones abroad. The problem is to be ready for<br />

opportunities in those few years when pest outbreaks overseas permit sizeable collections<br />

<strong>of</strong> candidate parasitoids. The manipulation <strong>of</strong> native parasitoids <strong>of</strong>fers two other avenues<br />

for action: inundative release <strong>of</strong> laboratory-reared stock, and management <strong>of</strong> forest<br />

habitats to favour the dynamics <strong>of</strong> parasitoid response to available hosts.<br />

Efforts to establish exotic parasitoids in spruce budworm populations <strong>of</strong> eastern<br />

<strong>Canada</strong> have not succeeded. In 1944-53, large numbers <strong>of</strong> four species (two tachinids,<br />

one sarcophagid, and one ichneumonoid) from unidentified western spruce budworms<br />

(British Columbia) were released at various sites from Manitoba to Newfoundland. In<br />

1948-56, 12 ichneumonoid species from European hosts, especially Choristoneura<br />

murinana (Hb.) (France, Germany, Czechoslovakia), were released in small numbers<br />

in northwestern Ontario and Quebec. No evidence <strong>of</strong> the establishment <strong>of</strong> any introduced<br />

species has been produced (McGugan & CoppeI1962), and there were no explanations<br />

<strong>of</strong> the causes <strong>of</strong> failure. Interest in further introductions was inhibited by these failures<br />

and by lack <strong>of</strong> sources <strong>of</strong> parasitoids abroad. However, during the 1960s unsuccessful<br />

efforts were made to obtain three ichneumonoids specific to C. murinana (Miller &<br />

Angus 1971).<br />

Interest was renewed when an outbreak <strong>of</strong> Choristoneura diversana (Hb.) on Japanese<br />

(todo) fir (Abies firma Sieb & Zucc.) in Hokkaido, Japan, occurred in the late 1960s.<br />

Collaboration among four institutes (Hokkaido Forest Experimental Station, Maritimes<br />

Forest Research Centre, Commonwealth Institute <strong>of</strong> Biological Control, and the Agriculture<br />

<strong>Canada</strong> Research Station at Belleville, Ontario) resulted in the shipment <strong>of</strong><br />

parasitoids by air from Japan to New Brunswick in 1970-75 (Tables 73 and 74). Further,<br />

a small collection <strong>of</strong> parasitoids from C. murinana in Czechoslovakia was shipped to New<br />

Brunswick in 1972. These introductions are detailed below.<br />

In the early 1970s, the theory that spruce budworm epidemics may start in isolated<br />

local epicentres led to interest in methods <strong>of</strong> quelling incipient outbreaks over small<br />

areas. At the Laurentian Forest Research Centre this interest developed into a research<br />

study <strong>of</strong> the native egg parasitoid, Trichogramma minutum Ril., as a candidate for<br />

inundative release, now briefly reported. As pressure to find a biological alternative to<br />

Open releases and recoveries <strong>of</strong> parasitoids against Choristoneura fumiferana (Clem.)<br />

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

Species and province Year Origin Number recovery<br />

Agria housei Shewell 1971 Ontario (laboratory 2800 1971<br />

New Brunswick culture) (14 adults)<br />

Nil in 1972-73<br />

Lissonota sp. 1973 Japan 59 Nil in 1974<br />

New Brunswick


CephaJogl)pta<br />

murinanae Bauer<br />

(Hymenoptera:<br />

Ichneumonidae)<br />

D. Testing <strong>of</strong> parasitoids 269<br />

success varying from 17 to 60% as indicated by dissection <strong>of</strong> samples. Substantial host<br />

larval mortality during early diapause was attributed partly to ovipositional wounding. A<br />

problem was to optimize survival <strong>of</strong> parasitized hosts by exposing candidate larvae long<br />

enough to achieve a high percentage <strong>of</strong> attacked hosts. but short enough to avoid a high<br />

percentage <strong>of</strong> superparasitism and fatal wounding. The exposure patterns usually resulted<br />

in a single egg per host; the egg hatched within a few weeks and the parasitoid overwintered<br />

as a first-instar larva.<br />

In rearing thousands <strong>of</strong> prospectively parasitized spruce budworrn larvae, only 70 C.<br />

laritis pre-pupae emerged from their hosts. Of these, 27 successfully pupated in a<br />

cocoon, a few adults emerged, and only one female oviposited. Efforts to rear its<br />

progeny were unsuccessful. Between autumn and late spring in all years, the percentage<br />

parasitism <strong>of</strong> sampled host stock declined sharply, almost to zero. Hypotheses to<br />

account for the decline included ovipositional injury to hosts, lack <strong>of</strong> cold hardiness in<br />

the parasitoid, host diet unfavourable to the parasitoid, encapsulation reaction, and<br />

other biochemical defence mechanisms.<br />

Mortality <strong>of</strong> hosts in hibernacula was 20-62%, higher than the usual 9-22% recorded<br />

from populations in their natural environment. Some <strong>of</strong> the additional mortality was<br />

attributed to ovipositional wounding, as C. laricis has a large ovipositor compared with its<br />

native homologue, Glypla fumiferanae (Vier.). However, mortality <strong>of</strong> hosts was not an<br />

important factor in reduction <strong>of</strong> percentage parasitism. It is unlikely that low winter<br />

temperature discriminated against parasitoids within their spruce budworm hosts. In a<br />

test for cold hardiness, overwintering larvae <strong>of</strong> C. fumiferana, G. fumiferanae, and C.<br />

larieis were rapidly supercooled down to -60°C. All three species showed ice-crystal<br />

formation in the -40 to -43°C range, which is below the coldest natural environment in<br />

New Brunswick. Thus the parasitoid stock <strong>of</strong> Hokkaido provenance appeared to be as<br />

hardy as the native stocks <strong>of</strong> parasitoids and hosts, just as one would expect from a<br />

comparison <strong>of</strong> the monthly mean temperatures <strong>of</strong> the two regions.<br />

When a stock <strong>of</strong> second-instar spruce budworrn larvae was held in storage from<br />

October to May at a constant temperature <strong>of</strong> 7°C, the rate <strong>of</strong> decline <strong>of</strong> parasitism was<br />

unusually rapid compared with stock held out-<strong>of</strong>-doors at much lower temperatures.<br />

Evidently the "cold room" was a selectively hostile environment to the parasitoid, but<br />

not to the host. This suggested that the decline in parasitism was temperature-related<br />

rather than time-related.<br />

Most <strong>of</strong> the loss <strong>of</strong> parasitism occurred in post-diapause stocks; however, no evidence<br />

implicating host diet appeared, as rearings with either artificial diet or with natural<br />

foliage produced healthy hosts and moribund or vanished parasitoids. Dissections<br />

revealed very few encapsulated parasitoid larvae, not enough to indicate a mechanism<br />

for decline in percentage parasitism. The expected parasitoids seemed to disappear<br />

without trace, but occasionally parasitoid head capsules, faintly yellowed and very<br />

membranous, were discovered within the body cavity <strong>of</strong> healthy maturing hosts. Thus<br />

the most plausible hypothesis is that the spruce budworrn is able to kill and absorb larvae<br />

<strong>of</strong> C. larieis, leaving the host unimpaired.<br />

A small shipment (Table 74) <strong>of</strong> pupae was air-freighted from Czechoslovakia to New<br />

Brunswick in July 1972. Some 1300 spruce budworrn larvae in hibemacula were exposed,<br />

and females oviposited vigorously with behaviour similar to C. larieis. A January<br />

sample <strong>of</strong> hosts wintered out-<strong>of</strong>-doors indicated 25% parasitism <strong>of</strong> survivors and 18%<br />

host mortality. Overwintered hosts were successfully reared in May on artificial diet, or<br />

on diet followed by foliage, but only four parasitoid pre-pupae matured, two dying<br />

within pupal hosts and two emerging from larval hosts but failing to pupate. The<br />

disappearance <strong>of</strong> the expected population <strong>of</strong> parasitoids is unexplained, but circumstances<br />

were similar to those <strong>of</strong> C. larieis.


270 1. W. Varty<br />

Lissonola sp.<br />

(Hymenoptera:<br />

Ichneumonidae)<br />

Trichogramma<br />

minutum Riley<br />

(Hymenoptera:<br />

Chalcidoidea)<br />

The biology <strong>of</strong> this ichneumonid species from C. diversana hosts on todo fir in Japan has<br />

been reported by Zw61fer (1972, personal communication). Shipments were air-freighted<br />

from Hokkaido to New Brunswick in 1972, 1973, and 1975 (Tables 73 and 74). Females<br />

mated soon after emergence from the cocoon, and were maintained on a diet <strong>of</strong> sugar,<br />

honey, and water for up to a month in insectary cages. No responses were observed to<br />

the presence <strong>of</strong> free-moving first-instar or spun-up second-instar spruce budworm.<br />

However, large samples dissected in winter revealed parasitism rates <strong>of</strong> 7% in 1972 and<br />

1% in 1973. No more than one egg shell was dissected from anyone host, although<br />

heavy superparasitism is the rule in C. diversana hosts. Although many thousands <strong>of</strong><br />

spruce budworm hosts were exposed, only one parasitoid emerged as a pre-pupa, and<br />

it failed to pupate. Three more mature larvae were dissected frot:T\ dead hosts, but no<br />

evidence <strong>of</strong> encapsulation or <strong>of</strong> tissue absorption <strong>of</strong> the parasitoids was found. Thus<br />

Lissonola sp. appears to be incompatible with spruce budworm because <strong>of</strong> low ovipositional<br />

stimulus to the parasitoid and some innate resistance by the host.<br />

The object <strong>of</strong>this study in Quebec in 1970-75 (F.W. Quednau 1974, personal communication,<br />

unpublished work reported to the Annual Review <strong>of</strong> Spruce Budworm Research,<br />

Canadian Forestry Service) was to explore the potential <strong>of</strong> this egg parasitoid for<br />

inundative release against rising spruce budworm populations in local epicentres. This<br />

native chalcidoid wasp is already present in all eastern North American forests, attacking<br />

eggs <strong>of</strong> scores <strong>of</strong> species, and typically contributing 5-25% parasitism <strong>of</strong> spruce budworm<br />

eggs. Apparently, scarcity <strong>of</strong> alternate hosts prevents the numerical response that T.<br />

m;nulUm might otherwise make to rising budworm density. Quednau believed that about<br />

2.5 x 10" female parasitoids per hectare would be needed for effective release in epicentres.<br />

The research problems facing the Laurentian Forest Research Centre were: (l)the<br />

maintenance <strong>of</strong> year-round research cultures <strong>of</strong> the parasitoids and suitable hosts;<br />

(2)comparison <strong>of</strong> commercial strains and arboreal ecotypes, and the selection <strong>of</strong> a<br />

type most responsive to laboratory rearing and the spruce budworm forest habitat;<br />

(3)determination <strong>of</strong> the parasitoid female's functional response to host density;<br />

(4)refinement <strong>of</strong> a mass-culture method, and (5)development <strong>of</strong> techniques for<br />

inundative release and assessment <strong>of</strong> efficacy.<br />

From 1970 to 1975 considerable biological knowledge and research experience was<br />

gained in the first three problems. Methods <strong>of</strong> air-freighting stock from the United States<br />

and Mexico were established, and techniques for small-scale rearing and storage <strong>of</strong><br />

parasitoids and host eggs were developed (Table 74). However, the application <strong>of</strong><br />

American commercial techniques <strong>of</strong> mass-rearing was thwarted by quarantine regulations,<br />

which prevented the utilization <strong>of</strong> the angoumois grain moth, Silotroga cerea/ella<br />

(Olivier), the most suitable host for Trichogramma cultures. Moreover, there were<br />

worker health problems (allergies) associated with mass-rearing.<br />

Progress was made in the selection <strong>of</strong> a Canadian ecotype well adapted to the forest<br />

environment and with superior host-finding potential. However, T. m;nulum was shown<br />

to have only a weak functional response to density <strong>of</strong> spruce budworm egg-masses; that<br />

is, in caged-tree experiments with varying parasitoid densities and host densities, the<br />

number <strong>of</strong> attacked host eggs per individual parasitoid was greater when egg-mass<br />

density was increased. However, increased density <strong>of</strong> parasitoids tended to lower<br />

success in finding hosts, perhaps because <strong>of</strong> mutual interference and increased superparasitism.<br />

Parasitism greater than 50% was not attained in cage experiments. Thus the<br />

prospects for useful impact on spruce budworm dynamics appear low, because only a<br />

drastic change in egg parasitism could have any significant effect on generation survival<br />

in spruce budworm outbreak densities (Morris 1963).<br />

The research was terminated in 1975 without definitive assessment <strong>of</strong> the feasibility <strong>of</strong><br />

application to spruce budworm control.


Agria housei Shewell<br />

(Diptera: Sareophagidae)<br />

Native parasitoids <strong>of</strong><br />

spruce budwonn larvae<br />

Recommendations<br />

D. Testing <strong>of</strong> parasitoids 271<br />

The native sarcophagid fly, Agria housei, has long been recorded as a pre-pupal and pupal<br />

parasitoid <strong>of</strong> spruce budworm and western spruce budworm, usually at low percentage<br />

parasitism. However, it merits consideration for inundative release because it is one <strong>of</strong><br />

the few budworm parasitoids that can be laboratory-produced in large quantities at<br />

acceptable costs; it can be reared directly on pork liver without an intermediate insect<br />

host. It has other characteristics that enhance its prospective application to the spruce<br />

budworm problem. However, the laboratory stock available for experimentation had<br />

been reared for many generations under artificial conditions, causing loss <strong>of</strong> cold hardiness<br />

and diapause induction (House 1967). Therefore there were doubts about the responsiveness<br />

<strong>of</strong> this stock to spruce budworm hosts and about its capacity to revert to univoltine<br />

reproduction.<br />

Field and laboratory tests were conducted in 1971 to gain experience in handling the<br />

species and to test its response to larval and pupal spruce budworm hosts (Tables 73 and<br />

74). The fly stock was provided by the Agriculture <strong>Canada</strong> Research Institute, Belleville,<br />

Ontario. Laboratory cage experiments designed to test hostlparasitoid densities failed<br />

to elicit any apparent interaction, and no parasitoid progeny issued. However, some 2800<br />

adult flies (34% female) were freely released into a balsam fir, Abies balsamea (L.) Mill.,<br />

stand highly infested with spruce budworm pupae, at Fundy National Park. From a<br />

sample <strong>of</strong> 1114 pupae collected within 10 days <strong>of</strong> release and within 7S m <strong>of</strong> release point,<br />

14 A. housei puparia were recovered. Based on estimates <strong>of</strong> the density <strong>of</strong> available<br />

hosts, this rate <strong>of</strong> recovery indicated an excellent response by the introduced stock. No<br />

further specimens were recovered from samples <strong>of</strong> host pupae in 1972 and 1973, but<br />

given the poor cold hardiness inheritance <strong>of</strong> the stock and the dispersal characteristics <strong>of</strong><br />

the species, detectable persistence at the release point was not expected. The potential<br />

for applied biological control under appropriate pre-outbreak conditions remains unassessed.<br />

The Canadian Forestry Service has maintained a sporadic surveillance <strong>of</strong> the effectiveness<br />

<strong>of</strong> native parasitoids attacking spruce budworm, especially G. fumiferanae and<br />

Apanteles fumiferanae Vier., which parasitize overwintering larvae. A study at the<br />

Maritimes Forest Research Centre was conducted from 1977 to 1980 on the premise that<br />

an understanding <strong>of</strong> the factors limiting parasitoid effectiveness might lead to methods<br />

<strong>of</strong> managing the forest habitat to stimulate parasitoid regulation <strong>of</strong> spruce budworm. The<br />

study was directed to the behavioural biology <strong>of</strong> adults <strong>of</strong> these two species (feeding<br />

habits, flight periodicity, distribution in the canopy), but did not reveal any clear avenue<br />

for practical application (R.S. Forbes 1980, unpublished work, Canadian Forestry<br />

Service).<br />

It is recommended that the Canadian Forestry Service (a) maintain contacts with<br />

CIBC concerning collection opportunities abroad; (b) investigate the compatibilities<br />

<strong>of</strong> a range <strong>of</strong> non-indigenous parasitoids with spruce budworm hosts; (c) develop<br />

rearing, release, and assessment skills and facilities; and (d) assess opportunities for<br />

incorporating parasitoid manipulation in integrated pest management programmes based<br />

on spruce budworm population dynamics.


272 I. W. Varty<br />

Literature Cited<br />

House, H.L. (1967) The decreasing occurrcnce <strong>of</strong> diapause in the fly Puudosarcophaga a!finis through laboratory-reared generations.<br />

Canadian Journal <strong>of</strong> Zoology 45,149-153.<br />

McGugan, B.M.; Coppel, H.C. (1962) Biological control <strong>of</strong> forest insects 1910-1958. In: A review <strong>of</strong> the biological control attempts against<br />

insects and weeds in <strong>Canada</strong>. Commonwealth Institute <strong>of</strong> Biological Control Technical Communication<br />

2,35-127.<br />

Miller, C.A.; Angus, T.A. (1971) Choristoneurafumiferana (Clemens), spruce budworm (lcpidoptera: Tortricidae). In: Biological control<br />

programmes against insects and weeds in <strong>Canada</strong> 1959- 1968. Commonwealth Institute <strong>of</strong> Biological<br />

Control Technical Communication 4,127-130.<br />

Minot, M.C.; Leonard, D.E. (1976) Host preference and development <strong>of</strong> the parasitoid Brachymeria intermedia in Lymantria dispar L..<br />

Galleria mellonella and Choristoneura fumiferana. Environmental Entomology 5.527-532.<br />

Morris, R.F. (1963) The analysis <strong>of</strong> generation survival in relation to age-interval survivals in the unspmyed areas. In: Morris, R.F. (Ed.)<br />

The dynamics <strong>of</strong> epidemic spruce bud worm populations. Memoirs <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> <strong>Canada</strong><br />

31,32-37.


274 I. W. Varty<br />

experimental biological insecticides with weak dispersion and self-sustenance in subsequent<br />

years. The Canadian Forestry Service (CFS) has undertaken fundamental and<br />

innovative research for more than 30 years, yet a cost-effective operational treatment for<br />

spruce budworm has not been achieved so far. In the 1970s, the emphasis was on<br />

isolation. mass-propagation, formulation, and spray-testing <strong>of</strong> baculoviruses. The economic<br />

problem was the high cost <strong>of</strong> virus propagation on diet-reared spruce budworm larvae,<br />

currently about $250 per sprayed hectare, or 30-40 times the cost <strong>of</strong> chemical treatment.<br />

The technical problem was the low rate <strong>of</strong> foliage protection, even when high larval<br />

mortality was attained.<br />

The probability <strong>of</strong> large-scale operational application in the 1980s is low. The future <strong>of</strong><br />

virus sprays may depend on development <strong>of</strong> a cheap culture method. Prospects for a selfspreading<br />

low-density inoculation method are poor. As no natural epizootic has been<br />

recorded, it appears that spruce budworm viruses have low natural infectiveness. Spray<br />

efficacy will probably demand high dosage and high density coverage <strong>of</strong> the target forest.<br />

Bacteria<br />

The commercial production <strong>of</strong> Bacillus thuringiensis Berliner (B.t.) and its Canadian<br />

registration for forest use have already been achieved. B.t. is widely used and effective<br />

for North American field crops, but has been generally less successful against forest<br />

pests in experimental trials. Its toxicology, formulation, aerial delivery techniques, and<br />

field assessment methods have been developed for use against spruce budworm. B.t. is<br />

essentially a biological insecticide, and is not designed to meet the self-regulating, selfdistributing,<br />

density-dependent criteria <strong>of</strong> c1assicial biological control. Canadian laboratory<br />

studies in the 1970s addressed the mode <strong>of</strong> action <strong>of</strong> the crystal, its structure, and assays<br />

<strong>of</strong> potency <strong>of</strong> field preparations; field research was designed to develop formulations,<br />

test aerial delivery techniques, and assess efficacy for population reduction and foliage<br />

protection. Even so, B.t. application is 4-5 times more costly and somewhat less<br />

reliable than corresponding treatments with the chemical insecticide fenitrothion. B.t.<br />

technology is still short <strong>of</strong> its full potential: potential future developments include<br />

selection <strong>of</strong> more potent strains, preparation <strong>of</strong> synthetic mimics <strong>of</strong> the crystal toxin,<br />

improvements in formulation, and more effective aerial delivery systems.<br />

Fungi<br />

The study <strong>of</strong> mycoses caused by entomogenous fungi holds a modest posItIon in<br />

research and development <strong>of</strong> biological methods for agricultural pests (Ferron 1978).<br />

International interest has focused mainly on the role <strong>of</strong> fungal epizootics in pest<br />

population regulation. but in practice few commercial products exist and experimentation<br />

in forestry, for example in the USSR, has been limited. The object <strong>of</strong> field application is<br />

to initiate an epizootic by artificial dissemination <strong>of</strong> inoculum, and research has been on<br />

the nature <strong>of</strong> infectiveness, methods <strong>of</strong> mass propagation, and methods <strong>of</strong>field distribution.<br />

Canadian research in this field, conducted mostly at the Forest Pest Management<br />

Institute and at Memorial University <strong>of</strong> Newfoundland, was exploratory in the 1970s and<br />

was on methods <strong>of</strong> culture <strong>of</strong> inoculum and the mode <strong>of</strong> spore formation and germination.<br />

Until 1980 no field spray experimentation for spruce budworm control had been attempted.<br />

Fungi are valuable agents for biological control because <strong>of</strong> their high specificity and high<br />

infective ness in the laboratory. However, their natural incidence in spruce bud worm<br />

populations is low, so the prospects for stimulating large-scale epizootics are correspondingly<br />

weak. Development <strong>of</strong> an operational cost-competitive treatment apears to be remote.


Protozoa<br />

E. Review <strong>of</strong> biological control opportunities 275<br />

Protozoan parasitism <strong>of</strong> agricultural pests is commonplace and is reported to reduce<br />

population vigour. Many species occur in forest defoliator insects, including two species<br />

<strong>of</strong> microsporidia in spruce budworm. However, in pest control history the applied use <strong>of</strong><br />

protozoa has been rare, because they are costly to propagate in quantity, and exhibit low<br />

infectiveness after artificial distribution.<br />

Canadian research on microsporidian parasites <strong>of</strong> spruce budworm was restricted to<br />

the Forest Pest Management Institute in the 1970s, which documented their natural<br />

incidence and parasitic role. Spray experiments on single trees demonstrated that<br />

parasitism rates can be artificially raised, but left the problem <strong>of</strong> mass production<br />

unsolved. Operational application is not yet practicable.<br />

Parasitoids<br />

<strong>Canada</strong> has enjoyed some success in the importation <strong>of</strong> exotic parasitoids to control<br />

introduced pest insects in both agriculture and forestry, but the record <strong>of</strong> releases to<br />

control native forest pests has been discouraging (Munroe 1971). However, it is not<br />

known whether the low success rate in the control <strong>of</strong> native hosts is associated with the<br />

separate evolutions <strong>of</strong> host and parasitoid, or stems from imperfect testing techniques.<br />

In agriculture, inundative release <strong>of</strong> a few species <strong>of</strong> natural enemies is routine practice,<br />

but experimentation against forest pests has been fragmentary and unsustained (Pschom­<br />

Walcher 1977).<br />

Canadian research on manipulation <strong>of</strong> spruce budworm parasitism was scant in the<br />

1970s, but was on a sounder base than the hit-and-miss releases <strong>of</strong> the 1940s and 1950s.<br />

Research was on field-cage studies <strong>of</strong> host and site compatibility with exotic parasitoids<br />

and studies <strong>of</strong> the numerical response <strong>of</strong> native parasitoids, with the inundative release<br />

method in mind.<br />

The case for increasing the Canadian effort with respect to spruce budworm parasitism<br />

is not strong if the probability <strong>of</strong> establishing additional natural enemies decreases with<br />

increasing diversity <strong>of</strong> the natural enemy complement. There is a need for better<br />

theoretical guidelines in order to achieve maximum success. The empirical approach -<br />

to collect homologous parasitoids abroad, then release and monitor their distribution<br />

and persistence - is unpredictable and does not increase understanding <strong>of</strong> the processes<br />

unless accompanied by studies <strong>of</strong> host-parasitoid compatibility.<br />

Predators<br />

In world agriculture, some striking successes have been achieved by inoculative release<br />

<strong>of</strong> exotic predaceous arthropods, but in Canadian forestry few such releases have been<br />

attempted (Munroe 1971). In general, the release <strong>of</strong> an exotic broad-spectrum predator<br />

gives rise to more concern about ecological integrity than the introduction <strong>of</strong> hostspecific<br />

parasitoids. In <strong>Canada</strong>, the only releases <strong>of</strong> non-native arthropod predators <strong>of</strong><br />

forest pests in 1969-80 were the successful introductions <strong>of</strong> two red wood ant species<br />

from Italy and Manitoba to Quebec (Finnegan 1978); the experiments show encouraging<br />

results in the control <strong>of</strong> spruce budworm populations in young stands. However, in<br />

general. the technical and economic feasibilities <strong>of</strong> predator releases against spruce<br />

budworm have not been assessed.<br />

Genetic approaches<br />

Pest insect genetics is an important means <strong>of</strong> biological control, because it leads to<br />

autocidal methods for manipulating popUlations. The technique <strong>of</strong> releasing sterile


276 I. W. Varty<br />

Further research<br />

Literature Cited<br />

males to interfere with the transmission <strong>of</strong> genes is best exemplified by the successful<br />

screwworm and fruit fly mntrol operations by the United States Department <strong>of</strong> Agriculture.<br />

During the 1970s many agencies around the world attempted pest control based on this<br />

principle (Whitten & Foster 1975).<br />

Exploratory research on spruce budworm genetics was conducted by CFS in the<br />

late 1970s, the only research agency with a declared interest. Such work is essentially<br />

long-term, demanding <strong>of</strong> innovation and risky as an investment; it includes investigations <strong>of</strong><br />

induced sterility, gene-damaging methods, and procedures for attaining population<br />

control. It may <strong>of</strong>fer opportunities for diminishing the budworm's reproductive success,<br />

but has potential mainly for small areas in the early stages <strong>of</strong> population increase.<br />

Prospects for early operational application are dim.<br />

Clearly none <strong>of</strong> the biological methods <strong>of</strong> control is ready to displace chemical spraying<br />

in a significant percentage <strong>of</strong> the area needing protection from spruce budworm.<br />

However, B.I. is efficacious and acceptable financially for small areas, especially in high<br />

value stands and in zones where chemicals are prohibited. The technical research, as<br />

recommended by Smirn<strong>of</strong>f & Morris in this chapter, should be strongly supported,<br />

because B.I. is close to being cost-competitive. It is the only non-chemical method able<br />

to replace the registered chemicals in the event <strong>of</strong> further restrictions, and it has<br />

considerable international research interest.<br />

For other entomopathogens, the main barrier to progress is the lack <strong>of</strong> methods for<br />

mass-propagation <strong>of</strong> agents at acceptable cost. Research agencies should re-examine<br />

the feasibility <strong>of</strong> hastening such capability, consulting with the many international<br />

agencies interested in similar, but problem-specific, developmental research. Because<br />

the search for new species or virulent strains <strong>of</strong> known entomopathogens has shown<br />

little promise, the prospects for induced genetic alteration should be monitored by close<br />

contact with scientific advances in agriculture. Meanwhile it is desirable to maintain the<br />

current steady progress reported already in this chapter.<br />

Research on parasitoids and predators is at low ebb. More basic research on hostparasitoid<br />

compatibility and on release and dispersion methods should be conducted.<br />

Research on genetic methods <strong>of</strong> autocidal control have been exploratory, but results<br />

have been sufficiently encouraging to justify further work.<br />

The contributors to each section <strong>of</strong> this chapter have recommended further long-term<br />

research. The potential for effective alternatives to chemical control is real, even if slight<br />

in most cases. Except for 8.1., none <strong>of</strong> these alternatives is likely to have a significant<br />

role in forest protection operations in the 1980s. However, because the spruce budworm<br />

is a long-term problem, and spruce-fir forest is a long-term resource, the need for an<br />

array <strong>of</strong> protection techniques for both remedial and preventive intervention is likely to<br />

persist.<br />

Beirne. B.P. (1975) Biological control attempts by introductions against pest insects in the field in <strong>Canada</strong>. Canadian Entomologist<br />

107.225-236.<br />

Ferron. P. (1978) Biological control <strong>of</strong> insect pests by entomogenous fungi. Annual Review <strong>of</strong> Entomology 23.409-442.<br />

Finnegan. R.J. (1978) Predation by Formica lugubris (Hymenoptera: Formicidae) on Choristoneura fumiferana (Lepidoptera: Tonricidae).<br />

Canadian Forestry Service Bi-monthly Research Notes. 34(1).3-4.<br />

Miller. C.A.; Vanyo I.W. (1975) Biological methods <strong>of</strong> spruce budworm control. Forestry Chronicle 51.16-19.<br />

Munroe. E.G. (1971) Status and potential <strong>of</strong> biological control in <strong>Canada</strong>. In: Biological control programmes against insects and weeds in<br />

<strong>Canada</strong>. 1959·68. Commonwealth Institule <strong>of</strong> Biological Control Technical Communication 4.213-255.<br />

Pschorn-Walchcr. H. (1977) Biological control <strong>of</strong> forest insects. Annual Review <strong>of</strong> Entomology 22.1-22.<br />

TInsley. T.W. (1979) The potential <strong>of</strong> insect pathogenic viruses as pesticidal agents. Annual Review <strong>of</strong> Entomology 24.63-87.<br />

Whitten. M.J.; Foster. G.G. (1975) Genetic methods <strong>of</strong> pest control. Annual Review <strong>of</strong> Entomology 20.461-476.


Pest Status<br />

Background<br />

Field Trials<br />

Chapter 48<br />

Choristoneura occidentalis Freeman,<br />

Western Spruce Budworm (Lepidoptera:<br />

Tortricidae)<br />

R.F. SHEPHERD and J.e. CUNNINGHAM<br />

Two outbreaks <strong>of</strong> the western spruce budworm, Choristoneura occidentalis Freeman,<br />

on Douglas fir, PseudolSuga menziesii (Mirb.) Franco, in British Columbia have been<br />

well documented and evidence <strong>of</strong> three more outbreaks during this century has been<br />

found by examination <strong>of</strong> tree radial sections. The current outbreak began in 1970 and<br />

still persists in 1980. It is more extensive and has continued longer than the previous<br />

outbreak, which lasted from 1954 to 1958.<br />

These outbreaks all occurred in the Frazer and Lillooet River valleys, which form an<br />

inter-zone between the moist coastal and dry interior zones <strong>of</strong> British Columbia. In the<br />

western and southern portions <strong>of</strong> these valleys, defoliation occurred in a mid·slope band<br />

but not in the valley bottoms. With time, populations have decreased in the western<br />

extremities <strong>of</strong> the infestation and increased toward the northeast, but during the current<br />

outbreak populations have decreased and then increased regionally more than once.<br />

Typical damage to trees includes reduction in height and radial growth, dieback, and<br />

deformity. Complete tree mortality has occurred in a few small patches but more<br />

commonly occurs in the understorey below larger heavily defoliated trees.<br />

The same viruses that infect C. occidentalis also infect spruce budworm, C. fumiferana<br />

(Clem.), and laboratory tests indicate that the former is more susceptible to a nuclear<br />

polyhedrosis virus (NPV) and a granulosis virus (GV) (Cunningham unpublished). In<br />

addition, populations <strong>of</strong> western spruce budworm are virtually free <strong>of</strong> the microsporidian<br />

parasites that are found at very high levels in older infestations <strong>of</strong> spruce budworm; it is<br />

possible that infection with microsporidia may be antagonistic to subsequent infection<br />

with viruses but this has to be verified.<br />

The first aerial spray trial using NPV on western spruce budworm was conducted in 1976<br />

on a 20.5·ha plot, using a fixed.wing aircraft fitted with boom and nozzle spray equipment.<br />

The dosage was 250 x 10" polyhedral inclusion bodies (PIB) per hectare; the aqueous<br />

formulation contained 25% molasses, 6% IMC 90-001 (International Minerals Corp.)<br />

UV protectant, and 1% Chevron® sticker; the emission rate was 9.4 I/ha and larvae<br />

were in the fourth, fifth, and sixth instars. Population reduction as a result <strong>of</strong> treatment,<br />

calculated by Abbott's formula (Abbott 1925), was 36%; the incidence <strong>of</strong> NPV in larvae,<br />

determined microscopically, was 6.8% in the treated plot and 1.6% in the check plot<br />

respectively. In 1977 the incidence <strong>of</strong> virus in the treated plot fell to 1.3% (Shepherd &<br />

Cunningham unpublished). This poor result was probably due to too Iowa dosage <strong>of</strong><br />

virus. applied too late.<br />

In 1978 the same equipment and emission rate were used to treat three 20-ha plots with<br />

NPV at a dosage <strong>of</strong> 750 x 10" PIBlha just after budflush when larvae were at the peak <strong>of</strong><br />

the fifth instar, with third, fourth. and sixth instars also present. The aqueous formulation<br />

277


27R R. F. Shcphcrd and J. C. Cunningham<br />

Evaluation <strong>of</strong> Control Attempts<br />

Recommendations<br />

contained 25% molasses, 3% Shade® (Sandoz Inc.) (the same product as IMC 90-001),<br />

and 0.05% Triton® B-1956 (Rohm and Haas Co. <strong>of</strong> <strong>Canada</strong> Ltd.) sticker. Fifteen days<br />

after spraying, NPV infection levels, detemined microscopically, were 55, 87, and 25%;<br />

and population reductions due to treatment were 0, 26, and 48% respectively. Successful<br />

adult emergence was lower in treated than in check plots,and the male:female ratio<br />

was 2: 1 in the treated plots compared to 1:1 in the check plots (Hodgkinson et al. 1979).<br />

In 1979 and 1980, follow-up studies were conducted on two <strong>of</strong> the three virus-treated<br />

plots, the third being abandoned because <strong>of</strong> a local population decline. In these two<br />

plots, NPV infection rates were recorded as 18 and 54% in 1979, and 7 and 17% in 1980.<br />

In 1979, population declines <strong>of</strong> 51 and 62% in the virus treated plots were recorded<br />

compared to increases <strong>of</strong> 46 and 201 % in the corresponding check plots. However, this<br />

trend was reversed in 1980; populations more than doubled in the virus-treated plots<br />

compared to 1979 levels and substantial population declines were noted in the check<br />

plots (Shepherd el al. 1982).<br />

Bacillus Ihuringiensis Berliner<br />

In 1978, four 4O-ha plots were treated with Thuricide® 16B (Sandoz Inc.), a commercial<br />

preparation <strong>of</strong> Bacillus Ihuringiensis Berliner (B.t.). The emission rate was 9.4 Vha and<br />

the dosage 20 x ur I. U ./ha. Fifteen days after spraying, populations were reduced by 32,<br />

66, 79, and 91% (Hodgkinson el al. 1979).<br />

In the plots treated with B.t., western spruce budworm populations increased in 1979<br />

in two <strong>of</strong> the four plots; in the third plot the population increased at a lesser rate and in<br />

the fourth it showed a small, but significant, decrease compared to corresponding check<br />

plots. In 1980 this trend was reversed, with untreated populations decreasing and B.t.treated<br />

populations increasing dramatically (Shepherd el al. 1982).<br />

Applications <strong>of</strong> NPV and B.I., in 1978 did not provide adequate control in that year, but<br />

NPV and possibly B.I. did provide some carry-over effect and continued to cause a<br />

population regulating effect a year later. This carry-over effect might have continued<br />

longer, but the plots treated in these trials were small and results were probably masked<br />

by immigration <strong>of</strong> moths. If either NPV or B.t. can regulate populations for a number <strong>of</strong><br />

years after a single application, their use may be justified economically.<br />

The same constraints that apply to the viruses on C. fumiferana also apply to C.<br />

occidentalis. However, the western species is virtually free from microsporidian<br />

parasites; preliminary laboratory tests have also shown that it is more susceptible to<br />

both NPV and av than microsporidia-free, laboratory-reared spruce budworm larvae.<br />

It therefore appears that western spruce budworm is a better candidate for population<br />

regulation with viruses.<br />

More field trials should be conducted to compare the efficacy <strong>of</strong> NPV and av.<br />

A test involving an isolated population, where moth immigration would be negligible,<br />

should be conducted in order to evaluate the long-term effects <strong>of</strong> these biological<br />

control agents.


Literature Cited<br />

CflOris(rmcllra occidelllalis Freeman. 279<br />

Abbott, W.S. (1925) A method <strong>of</strong> computing the erfeetiveness <strong>of</strong> an insecticide. Journal <strong>of</strong> Ecotromic Entomology 18.265-267.<br />

Hodgkinson, R.S.; Finnis. M.; Shepherd, R.F.; Cunningham, J.C. (1979) Aerial applications <strong>of</strong> nuclear polyhedrosis virus and Bacillus<br />

tllUTitlgietlsis against western spruce budworm. British Columbia Ministry <strong>of</strong> Forests/Canadian Forestry<br />

Service Joim Report 10. 19 pp.<br />

Shepherd, R.F.; Gray, T.G.; Cunningham, J.C. (1982) Effects <strong>of</strong> nuclear polyhedrosis virus nnd Dacilllts tllUritlgietlSis on western spruce<br />

budworm one nnd two xenrs after aerial application. Canarlian Entomologist 114,281-282.


Blank Page<br />

280


Pest Status<br />

Background<br />

Releases and Recoveries<br />

AgaIhis pumila (Ratz.)<br />

(Hymenoptera:<br />

Braconidae)<br />

Chapter 49<br />

Coleophora laricella (Hiibner), Larch<br />

Casebearer (Lepidoptera: Coleophoridae)<br />

I.S. OTVOS and F. W. QUEDNAU<br />

The larch casebearer. Coleophora laricella (Hubner). was accidentally introduced from<br />

Europe, probably on nursery stock. It was first recorded in North America in 1886 at<br />

Northampton, Massachusetts (Herrick 1912), and in Ottawa in 1905 (Retcher 1906).<br />

The casebearer spread rapidly and by 1947 infested most tamarack, Larix laricina (Du<br />

Roi) K. Koch, in Newfoundland. the Maritimes. and Ontario. spreading west as far as<br />

Thunder Bay (McGugan & Coppel 1962). By 1970 it was present in southeastern<br />

Manitoba. It was first found on western larch. Larix occidentalis Nutt .• around St.<br />

Maries, Idaho, in 1957 (Denton 1958). and in British Columbia in 1966 (Molnar et al.<br />

1966). The larch casebearer moved rapidly along major larch stands in the valleys, and<br />

became established throughout the range <strong>of</strong> western larch in southeastern British<br />

Columbia in varying degrees <strong>of</strong> infestation. Successive years <strong>of</strong> severe defoliation cause<br />

growth reduction. crown dieback. and subsequent tree mortality (McGugan & Coppel<br />

1962. Turnock et al. 1969).<br />

Successful control <strong>of</strong> the larch casebearer in eastern and central <strong>Canada</strong> (Webb &<br />

Quednau 1971) by the introduction <strong>of</strong> parasitoids and by similar biological control work<br />

in the northwestern United States (Denton 1972. Ryan & Denton 1973) led to a resumption<br />

<strong>of</strong> the use <strong>of</strong> parasitoids in 1974 to combat the larch case bearer in British Columbia.<br />

An intensive survey <strong>of</strong> larch casebearer parasitoids conducted in British Columbia in<br />

1973 yielded 32 species <strong>of</strong> Hymenoptera (Miller & Finlayson 1974). In Quebec, studies<br />

on Diadegma laricinellum (Strobl) were continued after 1968, but the project was<br />

terminated in 1974 because <strong>of</strong> low priority.<br />

Four species <strong>of</strong> parasitoids. Agathis pumila (Ratz.), Chrysocharis laricinellae (Ratz.).<br />

Diadegma laricinellum (Strobl). and Dicladocerus japonicus Yshm. have been released<br />

in southeastern British Columbia. and a few D. laricinellum were also liberated in<br />

Quebec. The details <strong>of</strong> these releases are summarized in Table 75.<br />

This species was first introduced into British Columbia in 1969 when parasitized larch<br />

casebearers. collected in Montana (USA) were placed in larch casebearer infested<br />

stands at Arrow Creek and Fruitvale. The species was recovered in 1972 at both sites.<br />

Between 1974 and 1977 more adults <strong>of</strong> A. pumila, this time from Europe. were released<br />

at other locations in the southeastern part <strong>of</strong> the province. In addition, larch casebearer<br />

larvae parasitized by A. pumila in the laboratory were placed in infested larch stands.<br />

and parasitoids from an old larch case bearer infestation in British Columbia were also<br />

relocated.<br />

2XI


282 I. S. Otvos and F. W. Quednau<br />

Chrys«haris<br />

buidneHae (Ratz.)<br />

(Hymenoptera:<br />

EuJophidae)<br />

Table 75<br />

The establishment <strong>of</strong> the parasitoids was monitored by rearing last-instar larch casebearer<br />

larvae and pupae collected at the release sites. These rearings showed that A.<br />

pumila became established, but its numbers seem to be declining as another parasitoid,<br />

C. laricinellae, is increasing over the years. Quednau (1970) has shown that the<br />

multivoltine C. laricinellae increased at the expense <strong>of</strong> A. pumila in Quebec.<br />

Summad' <strong>of</strong> parasitoid releases and relocations in British Columbia and Quebec between<br />

1969 an 1980<br />

No. and sex <strong>of</strong><br />

Species and locality Lat. Long. Origin parasitoids released Year<br />

Agathis pumiJa (Ratz.)<br />

Fruitvale 49.06"N 117.33o"V Montana 100 M & F 1969<br />

Arrow Creek 49.07"N 116.26"W Montana 100> M & F 1969<br />

Nelway 49.02"N 117.1nV Austria 87M+89F 1974<br />

Blewett 49.28"N 117.2SW France 15 M + 15 F 1974<br />

Creston 49.06"N 116.31o"V Italy 10 M + 24 F 1975<br />

East Arrow 49.09"N 116.26o"V Italy 2OM+32F 1975<br />

Thrums 49.23"N 1l7.33"W Italy 9M+32F 1975<br />

Pass Creek 49.25"N 117.36o"V Italy 17 M + 31 F 1975<br />

Blewett 49.28"N 117.25"W Italy 14 M + 31 F 1975<br />

North Salmo 49.13"N 1l7.l4"W 400 (parasitized larvae) 1976<br />

Christina Lake 49.02"N 118.10"w Austria 25M+30F 1976<br />

Ross Spur 49.11"N 117.28"W Austria 24M+28F 1976<br />

Rossland 49.04"N 1l7.47"W Italy & Austria 16 M + 36F 1977<br />

Cherryville SO.13"N 118.34"W (relocation from l00M& F 1978<br />

Arrow Creek)<br />

Chrysocharis laricinellae (Ratz.)<br />

Thrums 49.23"N 1l7.33"W Switzerland 6 M + 12 F 1974<br />

Shuttleworth Creek 49.1ffl 119.32"W (relocation from 512 M & F 1980<br />

B.C.)<br />

Diadegma laricinellum (Strobl)<br />

South Slocan 49.23"N 1l7.33"W Austria IOM+5F 1974<br />

Thrums 49.23 c N 1l7.33"W Austria 7 M + 5 F 1974<br />

East Arrow 49.09"N 116.26"W Italy 9M + SF 1975<br />

Thrums 49.23°N 1l7.33"W Italy 9M + SF 1975<br />

North Creston 49. 12°N 116.34"W Italy 12 M + 28 F + 25 1976<br />

parasitized larvae<br />

Joliette' 46.02"N 73.27"W own rearings 12 F (mated) 1972<br />

Dicladocerus japonicus Yshm.<br />

Blewett 49.28"N 1l7.25°W Japan 52 M + 215 F 1974<br />

Thrums 49.23°N 1l7.33"W Japan 5M+20F 1974<br />

• Located in Quebec; all others are in British Columbia .<br />

C. laricinellae was apparently introduced in the Pacific Northwest with its host or it was<br />

released inadvertently with the early introductions <strong>of</strong> A. pumila (Ryan et al. 1974). In<br />

1974 a small number <strong>of</strong> imported C. laricinellae were released in British Columbia; in<br />

1980 greater numbers <strong>of</strong> the parasitoid were relocated from older infestations in the<br />

province.


Dladegma I8ricineIJum<br />

(Strobl), formerly<br />

known as D. nBnB<br />

(Grav.), (Hymenoptera:<br />

Ichneumonldae)<br />

DlclBdocerus<br />

jBponlcus Yshrn.<br />

(Hymenoptera:<br />

Eulophidae)<br />

Evaluation <strong>of</strong> Control Attempts<br />

Recommendations<br />

Co/eop//Ora laricella (Hubner), 283<br />

Adults <strong>of</strong> this parasitoid were liberated in small numbers in British Columbia from 1974<br />

to 1976, but none <strong>of</strong> the species has been recovered. In Quebec, 12 mated females from<br />

laboratory rearings were released in 1972, but recoveries could not be made because the<br />

project was terminated. Quebec received small quantities <strong>of</strong> D. laricinellum from<br />

Austria between 1968 and 1973. Because the biology <strong>of</strong> D. laricinellum was incompletely<br />

known, laboratory studies on competition between it and A. pumila were conducted<br />

before it was released in the field. These studies have shown that host larvae attacked by<br />

A. pumila are normally avoided by D. laricinellum except at high densities <strong>of</strong> A. pumila,<br />

which are rare in nature. D. laricinellum proved to be intrinsically inferior and died from<br />

the presence <strong>of</strong> A. pumila when both parasitoid species occurred in the same host larva<br />

(Quednau, unpublished). Synchronous attacks on larch casebearer larvae by both<br />

parasitoid species are unlikely to occur in nature, because A. pumila attacks smaller<br />

needle-mining stages than D. laricinellum. Factors adversely affecting the sex ratio <strong>of</strong><br />

the progeny <strong>of</strong> this parasitoid were discussed by Ryan (1980). These factors comprise<br />

continuous supply with hosts, overmating, too-early contact with hosts, superparasitism<br />

and lack <strong>of</strong> sufficient time between attacks to encourage the fertilization <strong>of</strong> parasitoid<br />

eggs. Because some <strong>of</strong> these factors were not used to provide optimum conditions for<br />

the proper sex ratio during the laboratory rearings <strong>of</strong> D. laricinellum in Quebec, too<br />

few female parasitoids were obtained for field liberations (Quednau unpublished).<br />

This species was released in 1974 at Blewett, British Columbia, in relatively small<br />

numbers and has not been recovered to date.<br />

It is too early to evaluate the effect <strong>of</strong> the parasitoids on the larch casebearer in British<br />

Columbia. Data obtained so far indicate that both A. pumila and C. laricinellae have<br />

become established. The latter is now fairly common in this province (Otvos unpublished)<br />

and may have contributed to the reduction <strong>of</strong> the larch case bearer and hence <strong>of</strong> tree<br />

mortality.<br />

Further searches should be made for D. laricinellum; because <strong>of</strong> the failure to recover it,<br />

the role <strong>of</strong> this parasitoid in Canadian larch casebearer populations remains to be<br />

evaluated. Its presence in the ecosystem can be only beneficial because it also may serve<br />

as a host for C. laricinellae. The introduction <strong>of</strong> A. pumila has probably facilitated the<br />

distribution <strong>of</strong> C. laricinellae in British Columbia. Although no studies have been made<br />

on this subject, winter mortality <strong>of</strong> the parasitoids should be considerably less in British<br />

Columbia than in the eastern part <strong>of</strong> <strong>Canada</strong>, and the impact <strong>of</strong> C. laricinellae on larch<br />

case bearer populations should be stronger because <strong>of</strong> the more favourable climate in the<br />

west. Life-table studies, although time consuming, are essential to the understanding <strong>of</strong><br />

host-parasitoid population regulation mechanisms.


2M.. I. S. OtV()S and F. W. Quednau<br />

Literature Cited<br />

Denton. R.E. (1958) The larch c3scbearer in Idaho - a new defoliator for western forest. US Department <strong>of</strong> Agriculture Forest Service<br />

Research Note 51. Intermountain Forest and Range Experiment Station, Ogden, Utah, 6 pp.<br />

Denton. R.E. (1972) Establishment <strong>of</strong> Agathis pumila (Ratz.) for control <strong>of</strong>larch casebearer. and notes on native parasitism and predation in<br />

Idaho. US Department <strong>of</strong> Agriculture Forest Service Research Note INT·I64. 6 pp.<br />

Retcher. J. (1906) Insects injurous to Ontario crops in 1905. Forest and shade trees. <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> Ontario Report 36.89-90.<br />

Herrick. G.W. (1912) The larch cascbearer. Blllletin <strong>of</strong> the Cornell Agricultural Experiment Station No. 322.<br />

McGugan. B.M.; Coppel, H.C. (1962) Biological control <strong>of</strong> forest insects, 1910·1958. In: A review <strong>of</strong> biological control attempts against<br />

insects and weeds in <strong>Canada</strong>. Commonwealth Institute <strong>of</strong> Bioiogicill Contwl Technical Communication<br />

2.35-216.<br />

Miller. G.E.; Finlayson, T. (1974) Native parasitoids <strong>of</strong> the larch cascbearer, Coleophora laricella (Lepidoptera: Coleophoridae), in the west<br />

Kootenay area <strong>of</strong> British Columbia. Journal <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> British Columbia 74.14 - 21.<br />

Molnar, A.C.; Harris. J.W.E.; Ross, D.A. (1966) British Columbia Region. Annual Report <strong>of</strong> the Forestln.rect and Disease Survey, Ottawa<br />

1967. PI'. 108-123.<br />

Quednau. F. W. (1970) Competition and cooperation between Chrysoclwris laricinellae and Agathis pumila on larch casebearer in Quebec.<br />

Canadian Entomologist 102.602-612.<br />

Ryan. R.B. (1980) Rearing methods and biological notes for seven species <strong>of</strong> European and Japanese parasitoids <strong>of</strong> the larch casebearer<br />

(Lepidoptera: Coleophoridae). Canadian Entomologist 112.1239-1248.<br />

Ryan. R.B.; Denton, R.E. (1973) Initial releases <strong>of</strong> Chrysocharis laricinella and Dic/adocerus westwoodii for biological control <strong>of</strong> the larch<br />

casebearcr in the western United States. US Department <strong>of</strong> Agriculture Forest Sen-ice Research Note<br />

PNW·200. Pacific Northwest Forest and Range Experiment Station. Portland, Oregon. 4 pp.<br />

Ryan. R.B.; Bousfield, W.E.; Miller. G.E.; Finlayson, T. (1974) Presence <strong>of</strong> Chrysocharis Ia,icinellae. a parasite <strong>of</strong> the larch casebearer. in<br />

the Pacific Northwest. Journal <strong>of</strong> Economic Entomology 67.805.<br />

Tunnock, S.; Denton, R. E.; Carlson, C.E.; Janssen. W. W. (1969) Larch casebearer and other factors involved with deterioration <strong>of</strong> western<br />

larch stands in northern Idaho. US Department <strong>of</strong> Agriculture Forest Service Research Paper, INT-68.<br />

Intermountain Forest and Range Experiment Sation, Ogden, Utah. 10 pp.<br />

Webb. F.E.; Quednau, F.W. (1971) Coleophora laricella (Hiibner), larch casebearer (Lepidoptera: Colcophoridae) In: Biological control<br />

programme against insects and weeds in <strong>Canada</strong>, 1959·1968. Commonwealth Institute <strong>of</strong> Biological<br />

Contwl Technical Communication 4,131-143.


Pest status<br />

Chapter 50<br />

Coleophora serratelfa (L.), Birch Casebearer<br />

(Lepidoptera: Coleophoridae)<br />

A.G. RASKE<br />

The birch casebearer, Co/eophora semlIe/JQ (L.)I, is native to Europe; Salman (1929)<br />

postulated that the insect was accidentally introduced into northeastern North America<br />

about 1920. There it has become the most important defoliator <strong>of</strong> white birch, Betula<br />

papyri/era Marsh. It was discovered in Maine in 1927 (Salman 1929) and had spread to<br />

New Brunswick by 1933, to Nova Scotia by 1937 (Reeks 1951), to Ontario by 1944<br />

(Raizenne 1952), to Prince Edward Island by 1951, to Quebec by 1956, and to Manitoba<br />

by 1969 (Cochran 1974). It was discovered in western Newfoundland in 1953 and had<br />

spread throughout the island by 1971 (Raske 1974a). It was discovered in Victoria,<br />

British Columbia, in 1962 (probably from a separate introduction) and was causing<br />

severe defoliation <strong>of</strong> ornamental trees in about 100 ha <strong>of</strong> the city by 1980 (van Sickle<br />

personal communication). The pest has also been collected in Alberta (B. Wright 1981,<br />

personal communication). This casebearer thus appears to occur throughout <strong>Canada</strong><br />

(Fig. 11) but the exact distribution in central and western <strong>Canada</strong> is largely unknown.<br />

Identification records for Ontario were recently revised. Thus C. serratella has never<br />

been positively identified in Ontario but a thorough search has not been undertaken.<br />

The most common form <strong>of</strong> damage to trees is the browning and reduction <strong>of</strong> foliage<br />

caused by the feeding <strong>of</strong> late instar larvae in early summer (Bryant & Raske 1975). The<br />

larva feeds like a leafminer. It attaches its case to a leaf and mines as far as it can reach<br />

without leaving the case, creating somewhat rectangular mined patches. After a larva<br />

has eaten all the food within reach, it moves to another part <strong>of</strong> the leaf and repeats the<br />

process. The mined portion turns brown and this causes the scorched appearance <strong>of</strong><br />

severely damaged trees. The damage tends to become masked later in the summer by the<br />

continuous production <strong>of</strong> new foliage.<br />

The more severe form <strong>of</strong> damage occurs in early spring when large numbers <strong>of</strong> earlyinstar<br />

larvae destroy the flushing buds, causing twig, branch, and sometimes tree<br />

mortality (Bryant & Raske 1975); many insects starve.<br />

Outbreaks <strong>of</strong> the birch casebearer have been reported from Victoria, British Columbia,<br />

and from Quebec to Newfoundland. The insect probably reached St. John's, Newfoundland,<br />

in the late 1960s. It was extremely rare in Newfoundland in 1971 (Raske 19740) but<br />

had reached outbreak proportions by 1980. In a newly invaded area, populations tend to<br />

increase to a super-saturation level. After 1 or 2 years populations decline but continue<br />

to cause a high level <strong>of</strong> chronic defoliation <strong>of</strong> 20-50%. However, the post-outbreak<br />

population levels vary in both location and time.<br />

The life history and habits <strong>of</strong> the birch case bearer in the northeastern United States<br />

were reported by Salman (1929) and by Gillespie (1932). Guevremont & Juillet (1974)<br />

and Raske (1976) described aspects <strong>of</strong> the life history in <strong>Canada</strong>. Cosh an (1974) and<br />

Gepp (1975a) have summarized life history studies in Europe. The life cycle <strong>of</strong>the insect<br />

spans two calendar years. In <strong>Canada</strong>, eggs are laid in July on the underside <strong>of</strong> leaves.<br />

Larvae hatch in August and mine the leaf until early September. After moulting they<br />

cut a case from the leaf epidermis, crawl to a different area <strong>of</strong> the leaf, and feed until fall.<br />

I J.C. Bradley (In: Kloet, G.S.; Hinks, W.D. (Ells.) 1972 - Handbook for the identification<br />

<strong>of</strong> British insects, Vol. 2, Part 2. Lepidoptera. Royal <strong>Entomological</strong> <strong>Society</strong>,<br />

London) cites C. fuscedinella Zell. as a synonym <strong>of</strong> C. serralella (L.).<br />

285


Background<br />

Releases and Recoveries<br />

Coleophora serralella (L.). 287<br />

Just before leaf drop they crawl to branch crotches or to the base <strong>of</strong> leaf buds to<br />

overwinter. In the following spring, after moulting, they feed on developing leaves.<br />

When fully grown, by early summer, they pupate and adults emerge in July.<br />

Species <strong>of</strong> alder (Alnus spp.) are the main hosts in continental Europe (Gepp 1975a),<br />

but in the British Isles several species <strong>of</strong> birch are the most severely attacked trees<br />

(Coshan 1974). In North America (Guevremont & Juillet 1974, Bryant & Raske 1975)<br />

white birch is the most severely damaged species, although grey birch, Betula populi/olio<br />

Marsh., may also sustain continuous severe damage. The casebearer may complete its<br />

life cycle on several species <strong>of</strong> hardwood trees and shrubs, but in North America it is<br />

known to build up high population levels only on white and grey birch (Guevremont &<br />

Juillet 1974, Bryant & Raske 1975).<br />

In North America, the extent <strong>of</strong> mortality <strong>of</strong> various life stages has been determined in<br />

Quebec (Guevremont & Juillet 1974) and in Newfoundland (Raske unpublished). Life<br />

tables for the two regions differ substantially. In Quebec, mortality <strong>of</strong> first- and secondinstar<br />

larvae and pupal parasitism seem to be the most important mortality factors. In<br />

Newfoundland, highly variable egg mortality, from unknown causes, seems to regulate<br />

the population at a chronic high level. This egg mortality (Raske 1974b) has not been<br />

reported on mainland <strong>Canada</strong>. In fall, early instar larval mortality in Newfoundland is<br />

about 20%. Surviving larvae that overwinter suffer mortality <strong>of</strong> about 60% annually. In<br />

spring, larval mortality among those remaining is about 80% and pupal mortality is 25%.<br />

Total losses from parasitism <strong>of</strong> all stages is about 10%.<br />

The parasitoid complex <strong>of</strong> this casebearer has been studied both in Europe and in<br />

North America where biological control <strong>of</strong> this pest has been attempted. In its native<br />

Europe, parasitoids are more important in regulating the population <strong>of</strong> the birch<br />

casebearer (Pschom-Walcher 1969 to 1975, Gepp 1975b, 1975c, Coshan 1974) than in<br />

North America (Guevremont & Juillet 1975, Raske 1978). However, the parasitoid<br />

complexes on the two continents resemble one another. Many genera are native to both<br />

regions: Scambus, ltoplectis, Gelis, Orgilus, Agathis, Habrocytus, Cirrospilus, and<br />

Chrysocharis. The species on each continent differ, except for Habrocytus semotus<br />

(Walker). Apanteles spp. and Campoplex spp. are dominant parasitoids <strong>of</strong> this casebearer<br />

in Europe, but are virtually absent on it in North America. The main difference between<br />

Quebec and Newfoundland is that Agathis cincta was common in Quebec and absent in<br />

Newfoundland.<br />

Most parasitoid species in North America are parasitic on many lepidopterous insects<br />

and are best considered incidental parasitoids that have adapted to this host.<br />

In 1968 the Canadian Forestry Service, in co-operation with the aBC, initiated a<br />

biological control programme against the birch casebearer in Newfoundland; introductions<br />

<strong>of</strong> European parasitoids began in 1971 and terminated in 1975 (Raske 1977). Two<br />

species complexes <strong>of</strong> parasitoids were released: Campoplex (= Porizon) spp. (Hymenoptera:<br />

Ichneumonidae) and Apanteles spp. (Hymenoptera: Braoonidae) (Table 76). The Campoplex<br />

spp. consisted <strong>of</strong> C. borealis (Zett.) and an undescribed species. The Apanteles complex<br />

included three species: predominantly A. coleophorae (Wilk.) and a few each <strong>of</strong> A.<br />

mesoxanthus Ruschka and A. corvinus Reinh. Living individuals <strong>of</strong> neither complex


Blank Page<br />

290


Pest Status<br />

Background<br />

Chapter 51<br />

Fenusa pusilla (Lepeletier), Birch Leafminer<br />

(Hymenoptera: Tenthredinidae)<br />

F.W. QUEDNAU<br />

Fenusa pusilla (Lepeletier) is an important endemic pest on ornamental birches, Betula<br />

spp., in <strong>Canada</strong>. This insect was accidentally introduced from Europe into the State <strong>of</strong><br />

Connecticut, USA, in 1923 (Friend 1933). The tree species attacked are Betula papyri/era<br />

Marsh., B. populi/olia Marsh .• B. lema L.. and the European B. verrucosa Ehrh. and its<br />

varieties. The present distribution <strong>of</strong> the sawfly in <strong>Canada</strong> is mainly from Newfoundland<br />

to Thunder Bay in northwest Ontario. but it has recently been recorded from the Prairie<br />

Provinces (Fig 12). The infestation level <strong>of</strong> F. pusilla in <strong>Canada</strong> is usually high and<br />

remains relatively constant from one year to another. The insect, by its oviposition and<br />

the ensuing feeding <strong>of</strong> larvae in leaf mines, causes discolouration <strong>of</strong> the foliage by the<br />

destruction <strong>of</strong> parenchyma cells. The affected trees have an unsightly brownish discolouration<br />

<strong>of</strong> the leaves. which drop prematurely. Repeated infestations <strong>of</strong> the sawfly<br />

weaken a tree and make it prone to attack by other insects or diseases. There are about<br />

three overlapping generations <strong>of</strong> F. pusilla in a year. The biology <strong>of</strong> this insect in Quebec<br />

has been described by Cheng & leRoux (1965) and Guevremont (1975).<br />

A preliminary study <strong>of</strong> biological control agents affecting F. pusilla in Europe was<br />

undertaken by the Commonwealth Institute <strong>of</strong> Biological Control (CIBC) at Delemont.<br />

Switzerland (Pschorn-Walcher & Eichhorn 1968). The observation that F. pusilla is<br />

much less abundant in Central Europe than in <strong>Canada</strong> led to the decision by the<br />

Canadian Forestry Service to introduce parasitoids. In 1972 the Newfoundland Forest<br />

Research Centre initiated a programme at Pasadena, Newfoundland. In 1974 the project<br />

was transferred for technical reasons to the Laurentian Forest Research Centre at Ste­<br />

Foy, Quebec. The aim <strong>of</strong> this biological control programme was the implantation <strong>of</strong> the<br />

two hymenopterous parasitoid species, Latlrrolestes nigricol/is Thoms. and Grypocentrus<br />

albipes Ruthe. For information on the techniques employed for the releases and recoveries<br />

the reader is referred to reports by Guevremont & Quednau (1976, 1978).<br />

Releases and Recoveries The numbers <strong>of</strong> parasitoids released in <strong>Canada</strong> are listed in Table 77.<br />

LaIhroIestes nigrirolJis<br />

Thoms. (syn.<br />

Priopoda nigricollis<br />

(Thoms.»<br />

(Hymenoptera:<br />

(chneumonidae)<br />

This is a multivoltine solitary endoparasitoid <strong>of</strong> semi-mature and mature larvae <strong>of</strong> the<br />

leafminer. Its biology in Europe was described by Eichhorn & Pschorn-Walcher<br />

(1973). Additional observations on its mating and oviposition behaviour were made by<br />

Quednau & Guevremont (1975). It is the most important natural enemy <strong>of</strong> F. pusilla<br />

in Central Europe. It is probably monophagous and well synchronized with successively<br />

overlapping stages <strong>of</strong> the host in the field. In captivity the female parasitoids are easily<br />

mated. Partial encapsulation <strong>of</strong> parasitoid eggs in the host larva, the incidence <strong>of</strong><br />

superparasitism. multiple parasitism by other parasitoid species. and entry in diapause<br />

291


Pest Status<br />

Chapter 52<br />

Gilpinia hercyniae (Hartig), European<br />

Spruce Sawfly (Hymenoptera: Diprionidae)<br />

L.P. MAGASI and P.O. SYME<br />

The European spruce sawfly, Gilpinia hercyniae (Hartig), a major pest <strong>of</strong> spruce (Picea<br />

spp.) in eastern <strong>Canada</strong> during the 1930s and early 19405, has lost its economic<br />

importance with the collapse <strong>of</strong> outbreaks in the Gaspe Peninsula and in northern New<br />

Brunswick (Neilson er aI. 1971). Populations have remained generally low since the<br />

early 1940s. The insect, and specifically its biological control by parasitoids and<br />

pathogens, has thus received very little attention since the last review in 1969. Consequently,<br />

the information presented here is meagre. It deals mainly with major population<br />

changes and changes in distribution <strong>of</strong> the insect during the past decade. No research<br />

has been conducted on the biological control <strong>of</strong> the spruce sawfly in <strong>Canada</strong> since 1969<br />

and few references are available. Distribution <strong>of</strong> the pest is summarized in Fig. 13.<br />

Newfoundland<br />

Populations were generally low, with occasional minor increases. Larval numbers in<br />

western Newfoundland increased from l.4/tree in 1970 to 18.0/tree in 1971 and to<br />

24.OItree in 1972. However, by 1974 the population decreased to 1.0 larva/tree throughout<br />

the province and has remained at that level.<br />

The nuclear polyhedrosis virus (NPV) Borrelinavirus hercyniae, introduced in 1943,<br />

is assumed to be a principal factor in controlling outbreaks but the introduction <strong>of</strong><br />

invertebrate parasitoids and perhaps even the masked shrew, Sorex cinereus cinereus<br />

Kerr, have doubtless improved the biological control <strong>of</strong> this insect.<br />

Maritimes<br />

Population levels were generally low in the region throughout the decade. The highest<br />

number recorded was in Pictou County, Nova Scotia, where 17 larvae were found on<br />

four white spruce trees, Picea glauca (Moench) Voss, in 1978. The NPV B. hercyniae<br />

was last found in larvae in 1975. The dipterous parasitoid Drino bohemica (Mesn.) and<br />

the hymenopterous parasitoid Exenterus veIlicatus Cush. were last reared in 1977, but it<br />

should be added that, apart from these rearings, no efforts have been made to identify<br />

these control agents.<br />

Neilson et aI. (1971) in an earlier edition <strong>of</strong> this book described an increase in the<br />

sawfly population after the cessation <strong>of</strong> chemical insecticide spraying to control spruce<br />

budworm, Choristoneurafumiferana (Clem.). This situation occurred again in the early<br />

1970s when the monitoring plot was sprayed for 2 years. Sawfly levels increased<br />

immediately afterwards and remained high for a number <strong>of</strong> generations until, presumably,<br />

the host and its control agents re-established their balance.<br />

Quebec<br />

The insect was present in low numbers throughout its known range, with minor<br />

fluctuations. Laboratory rearings showed the presence <strong>of</strong> the NPV B. hercyniae in 1971,<br />

1973, 1974, and 1975, and the virus is considered a major factor in controlling the<br />

population.<br />

295


Ontario<br />

Gilpitlia Ill'rcylliae (Hartig), 297<br />

The insect has been sporadically reported, and occurs throughout the province within<br />

its known range but usually in very low numbers. In 1977, light defoliation <strong>of</strong> white<br />

spruce occurred in parts <strong>of</strong> southwestern Ontario.<br />

The Forest Insect and Disease Survey participated in a parasitoid identification<br />

project with CIBC in 1973. Of 160 sawfly cocoons, 7 adults <strong>of</strong> E. vellicatus emerged<br />

(4.4% parasitism); the parasitoids were mostly male.<br />

Manitoba<br />

Evaluation <strong>of</strong> Control Attempts<br />

Literature Cited<br />

The insect was first reported in Manitoba in 1969, when it was found in low<br />

populations on 5200 km! adjacent to the Ontario border. In 1970, populations were<br />

reported on a further 2600 kml, as far north as Otter Falls and Meditation Lake in the<br />

Whiteshell Provincial Park, and as far west as the Agassiz and the Sandilands Provincial<br />

Forests. It has not extended its range since then and population levels have remained<br />

low.<br />

The European spruce sawfly appears to be controlled, largely by biological agents.<br />

However, there may be differences in areas that are regularly subjected to chemical<br />

sprays. Neilson et al. (1971) stated that when spraying stopped, there was a sharp<br />

increase in sawfly populations - at least until the balance between host and control<br />

agent was re-established. This phenomenon was observed again in the 19705. In each<br />

case, the spray period lasted only 2-4 years. It is not known what is happening in areas<br />

<strong>of</strong> sustained chemical insecticide application: the European spruce sawfly, obviously,<br />

is kept at low populations but possibly not by biological control agents. Nor is it known<br />

what the status <strong>of</strong> these agents is or if they would rapidly re-establish the balance<br />

necessary for control once chemical insecticide spraying ceased. High-value spruce<br />

plantations, which were practically non-existent in the 1930s and 194Os, are now<br />

plentiful and provide a bountiful food supply for the sawfly.<br />

Neilson, M.M.; Martinez. R.; Rose, A.H. (1971) Diprion hercyniae (Hartig), European spruce sawfly. In: Biological control programmes<br />

against insects and weeds in <strong>Canada</strong>. 1959-1968. Commonweahh /outilUle <strong>of</strong> Biological Control Technical<br />

Communication 4,136-143.


Blank Page<br />

298


-(I)<br />

C<br />

c:::<br />

o<br />

'';:;<br />

::::l<br />

.c<br />

';:<br />

300 L. P. Magasi and O. A. Van Sickle


Leucoma salids (L.), 301<br />

was recorded at Upper Ferry and light damage occurred on willow in the Stephenville<br />

area. In 1980 the outbreak spread from St. Andrews to Fischell's River, mostly along the<br />

major rivers where natural stands <strong>of</strong> balsam poplar occur.<br />

Maritimes<br />

In 1969, over 3200 ha <strong>of</strong> woodland trees, trembling aspen, Populus tremuloides Michx.,<br />

and largetooth aspen, Populus grandidentata Michx., sustained moderate to severe<br />

defoliation in southeastern New Brunswick. The outbreak began in 1967, persisted for 3<br />

years, and collapsed in 1970. The collapse was partly due to biological control agents. At<br />

Elgin, New Brunswick, cocoons <strong>of</strong> A. solitarius were numerous on tree trunks in the<br />

winter <strong>of</strong> 1969-70 and counts <strong>of</strong> up to 732 cocoonslm l <strong>of</strong> bark surface were made. A<br />

cytoplasmic polyhedrosis virus, known in the maritimes since 1954, infected up to 67%<br />

<strong>of</strong> field-collected laboratory-reared larvae in 1968 near Anagance, within the same area<br />

as the L. salids infestation.<br />

Small outbreaks occurred in natural stands in 1974 in south-central New Brunswick,<br />

in 1976 in the central and in 1979 in the east-central parts <strong>of</strong> the province. All these<br />

outbreaks affected trembling aspen, but in 1974 largetooth aspen was also affected.<br />

None <strong>of</strong> the outbreaks lasted most than a single season.<br />

The insect was found on ornamental silver and Carolina poplars Populus alba L. and<br />

P. canadensis Moench, and willow, during 1969-80 at locations scattered throughout<br />

the region, resulting in varying degrees <strong>of</strong> defoliation.<br />

The parasitoids reared from satin moth in 1969-70 and their frequency are listed as<br />

follows:<br />

Apanteles solitarius (Ratz.)<br />

Compsilura concinnata (Mg.)<br />

12 locations; mean parasitism 51%<br />

(range 16-100%)<br />

6 locations; mean parasitism 15%<br />

(range 5-55%)<br />

Exorista sp. 1 location; percentage parasitism unknown<br />

The mean parasitism from all rearings was 28% for A. solitarius and 4% for C.<br />

concinnata. No rearings or insect analyses have been done since 1970.<br />

Quebec<br />

The general distribution <strong>of</strong> satin moth has increased considerably during the past decade<br />

and now includes all eastern Quebec and both sides <strong>of</strong> the St. Lawrence River. Infestations,<br />

causing defoliation that varied from trace to severe, were localized and were reported<br />

every year with the exception <strong>of</strong> 1970. No examinations <strong>of</strong> parasitism were made from<br />

1969-80.<br />

Ontario<br />

The insect was first reported in Ontario in 1972 from Cornwall and Lancaster Township,<br />

where it caused severe defoliation <strong>of</strong> individual silver poplar trees. By 1976 it had spread<br />

to near Ottawa. By 1980 it had reached west <strong>of</strong> Brockville and the severity <strong>of</strong> attack was<br />

reported on the increase. There are no records <strong>of</strong> parasitoids from the province. but<br />

again no check was made.


302 L. P. Magasi and G. A. Van Sickle<br />

Evaluation or Control Attempts<br />

Literature Cited<br />

British Columbia<br />

The general distribution has not changed appreciably during the reporting period but<br />

satin moth infestations were recorded in three new areas; at Revelstoke in 1972, at Avola<br />

in the North Thompson Valley in 1975, and at Rossland in 1976. Whether these represent<br />

a true range extension or are only outlying areas cannot be stated with certainty because<br />

<strong>of</strong> the discontinuous nature <strong>of</strong> the tree-host and insufficient collections <strong>of</strong> insects.<br />

Infestations usually occurred on small groups <strong>of</strong> a few shade trees. However, starting in<br />

1973 a few severe outbreaks on trembling aspen and black cottonwood, Populus<br />

trichocarpa Torr. & Gray, were observed near Merritt, and by 1976 almost 1500 ha <strong>of</strong><br />

natural stands were severely defoliated between Aspen Grove and Kamloops. Another<br />

outbreak, near Rossland, occurred in 1977 on over 140 ha <strong>of</strong> trembling aspen and small<br />

patches <strong>of</strong> trees were again defoliated in both these areas in 1978.<br />

In contrast to 67 rearing records during 1959-68, there were only two during 1969-80: in<br />

1972, 3 larvae reared from Revelstoke yielded one native parasitoid, Exorisla mella<br />

(Walker), and in 1976, 24 larvae, collected near Victoria, were reared: 6 were parasitized<br />

and 11 Apanteles melanoscellus (Ratt.) were recovered.<br />

Field observations recorded, but apparently unsubstantiated by collections, include:<br />

(l)in 1969, larvae were heavily parasitized in Lewis Park, Courtenay, where some<br />

trembling aspen was 80% defoliated; (2)in 1969, cocoons <strong>of</strong> Apanleies spp. were<br />

numerous on the trunks <strong>of</strong> severely defoliated black cottonwood at Slocan City; (3)in<br />

1975, A. solilarius was present in the second and last year <strong>of</strong> defoliation <strong>of</strong> black<br />

cottonwood over 1 ha along Birkenhead Lake.<br />

In a collection <strong>of</strong> more than 50 larvae from Lillooet in 1975, 6 were infected with a<br />

Beauveria sp. <strong>of</strong> fungus.<br />

The information presented here deals mainly with major population changes and changes in<br />

distribution <strong>of</strong> the host insect during the past decade; there are few instances where<br />

reference to biological control is made. No research has been conducted on the biological<br />

control <strong>of</strong> satin moth in <strong>Canada</strong> during the review period and none is recommended,<br />

given the minor economic importance <strong>of</strong> the pest. We have no information to alter the<br />

assessment given in the previous review, that parasitoids must recieve major credit for<br />

helping to shorten periods <strong>of</strong> pest outbreak and thus for reducing damage (Forbes & Ross<br />

1971).<br />

Forbes, R.S.; Ross, D.A. (1971) Stilnoptio salicis (L.). satin moth. In: Biological control programmes against insects and weeds in <strong>Canada</strong>,<br />

1959-1968. Commonwtalth lnslilutt <strong>of</strong> Biological Control Technical CommUllication 4,205-212.


Pest Status<br />

Background<br />

Chapter 54<br />

Lymantria dispar (L.), Gypsy Moth<br />

(Lepidoptera: Lymantriidae)<br />

K.J. GRIFFITHS and F.W. QUEDNAU<br />

The natural distribution <strong>of</strong> the gypsy moth, Lymantrio dispar (L.), extends across the<br />

whole <strong>of</strong> Europe, including Norway and Sweden up to about 58"N and all the temperate<br />

areas <strong>of</strong> North Africa and Asia, including Japan. Outbreaks occur periodicaUy throughout<br />

most <strong>of</strong> this range (Leonard 1974). The gypsy moth was accidentally introduced into<br />

North America at Medford. Massachusetts, in 1869, and within 20 years had become a<br />

serious local pest that required control (Burgess 1914). This pest has continued to cause<br />

serious periodic damage in spite <strong>of</strong> almost continuous work to control it. It is now found<br />

in the United States throughout the New England States, New York, New Jersey, and<br />

Pennsylvania, and has also been recorded in Delaware, Maryland, Virginia, and West<br />

Virginia. Isolated infestations have been reported in Washington, Oregon, California,<br />

Michigan, Wisconsin, Illinois, and Ohio (Forest Pest Survey Report, Virginia Division<br />

<strong>of</strong> Forestry, Sept.-Dec., 1980, unpublished report).<br />

The gypsy moth was first recorded in <strong>Canada</strong> in 1924 in Stanstead and St-Jean<br />

Counties, Quebec (McLaine 1925). A second invasion occurred in New Brunswick in<br />

1936 (McLaine 1938). Both these invasions apparently died out (Brown 1968) but a third,<br />

into Quebec south <strong>of</strong> Montreal in 1959 (Brown 1968), did not disappear and the gypsy<br />

moth has continued to spread in all directions from that area. The first invasion into<br />

Ontario occurred on Wolfe Island, just south <strong>of</strong> Kingston in 1969 (Sippell et 01. 1970),<br />

and it was reported on nearby Howe Island and the mainland near Kingston in 1970<br />

(Sippell et al. 1971). The Plant Quarantine Division <strong>of</strong> Agriculture <strong>Canada</strong> reports that<br />

the 1979 distribution <strong>of</strong> gypsy moth extends from approximately 60 km west <strong>of</strong> Quebec<br />

City to Morrisburg, Ontario, and north to Ottawa, and from near Brockville, Ontario, to<br />

25 km west <strong>of</strong> Belleville, with a northward extension <strong>of</strong> approximately 70 km. There is<br />

also a population in Greater Toronto and in Oakville, Ontario (Fig. 15). A small<br />

infestation was detected over a two-block area in Vancouver, British Columbia, in the<br />

summer <strong>of</strong> 1978, and a control operation was undertaken in 1979. There were no male<br />

adult recoveries in pheromone traps set out in the area in the fall <strong>of</strong> 1979 and the<br />

infestation is believed to have been averted. On the east coast there seems to be an<br />

annual influx <strong>of</strong> male moths from Maine. In Nova Scotia and New Brunswick males<br />

have been recovered in pheromone traps every year since 1971, with a marked increase<br />

in both provinces in the past 4 years. There were no male recoveries on Prince Edward<br />

Island in the first 6 years <strong>of</strong>trapping, but in 1979 nearly 47% and in 1980,19% <strong>of</strong> the traps<br />

produced males. However. egg masses have never been recovered in the maritimes in<br />

this period (Magasi 1981).<br />

Since its arrival in North America, the gypsy moth has acquired a large complex <strong>of</strong><br />

native insect parasites, and predators. Griffiths (1976) lists 22 species <strong>of</strong> insect parasites<br />

and Sabrosky & Reardon (1976) added 4 more to this list. In addition there are 17 species<br />

<strong>of</strong> insect predators, 4 species <strong>of</strong> mammalian predators (Griffiths 1976), and 46 species <strong>of</strong><br />

303


Lymcltllria dispar (L.). 305<br />

avian predators (McAtee 1911) recorded from the United States. A major programme <strong>of</strong><br />

importation and liberation <strong>of</strong> exotic insect parasites and predators was carried on in the<br />

United States from 1905 to 1933 (Dowden 1962) in which 44 species <strong>of</strong> insect parasites<br />

and 9 species <strong>of</strong> predators were released. Thirteen insect parasites and one predator<br />

became established. As pointed out by Ticehurst et 01. (1978) these natural enemies have<br />

not prevented the dispersal <strong>of</strong>the gypsy moth in North America, but their influence may<br />

be responsible for initiating population collapses and for maintaining relatively stable<br />

infestations in many areas following a collapse. The introduction programme was<br />

revived in the early 1970s and is continuing. Eight species <strong>of</strong> insect parasite and one<br />

species <strong>of</strong> nematode have been released. Four <strong>of</strong> the insect species were released<br />

unsuccessfully in earlier work (Griffiths 1976). None <strong>of</strong> these nine species is known to be<br />

established.<br />

Fourteen <strong>of</strong> the 26 species <strong>of</strong> native parasitoids that attack the gypsy moth in the United<br />

States have been recovered from other hosts in areas <strong>of</strong> Ontario and Quebec adjacent to<br />

the American infestation. Also 11 <strong>of</strong> the 17 insect predators, all <strong>of</strong> the four mammalian<br />

predators and 43 <strong>of</strong> the 46 species <strong>of</strong> avian predators are known to be in this area <strong>of</strong><br />

<strong>Canada</strong> (Griffiths 1976). In addition, four exotic parasitoids established in the United<br />

States were present in southern Ontario and Quebec before the gypsy moth was recorded.<br />

Three <strong>of</strong> these species, Compsilura concinnata (Meigen), Apanteles lacteicolor Vier., and<br />

Meteor/IS versicolor (Wesm.) had been introduced against the browntail moth, Euproctis<br />

chrysorrhoea L., and the satin moth, Leucoma salicis L.. and one, Exorista larvarum (L.).<br />

apparently dispersed naturally from the United States.<br />

Several studies have been carried out recently to determine which parasitoid species<br />

are attacking the gypsy moth in <strong>Canada</strong>. The first <strong>of</strong> these investigations, in 1974 and<br />

1975, involved the collection and rearing <strong>of</strong> gypsy moth larvae from four woodlots near<br />

Kingston and three woodlots near Cornwall, Ontario. From these rearings, one native<br />

and three introduced parasitoid species were recovered (Table 79) (Griffiths 1977). The<br />

second study was carried out in 1977 and 1978 near Havelock, south <strong>of</strong> Montreal. and<br />

Mt-St-Hilaire, east <strong>of</strong> Montreal, Quebec. Larvae and pupae were collected and reared,<br />

and one more native and four more exotic species recovered (Table 79) (Madrid &<br />

Stewart 1980). A third study, carried out in 1978-80, used Malaise traps to capture<br />

adults at Mt-St-Hilaire and Mt-St-Bruno. east <strong>of</strong> Montreal. In this survey five additional<br />

native species were found (Table 79). The final study, carried out in 1979 and 1980, was a<br />

repetition <strong>of</strong> the work done by Griffiths (1977); two new native species and one exotic<br />

species, Ooencyrtus kuvanae (How.). were recovered (Table 79).<br />

Howard & Fiske (1911) noted that naturally occurring pathogens were playing a role<br />

in gypsy moth population regulation in the United States shortly after the introduction <strong>of</strong><br />

the insect there. Much work has been done on pathogens since then. Bacterial and<br />

fungal pathogens are a minor influence (Doane 1971, Podgwaite & Campbell 1972,<br />

Majchrowicz & Yendol 1973), but viral pathogens, especially the nuclear polyhedrosis<br />

virus (NPV) Borrelinavirus reprimens Holmes, are known to be a major factor in<br />

reducing outbreaks (Doane 1970, Campbell & Podgwaite 1971). The gypsy moth NPV<br />

was registered for use in the United States in 1978, and quantities <strong>of</strong> it have since been<br />

produced and marketed as Gypcheki!!) (US Department <strong>of</strong> Agriculture) (Doane & McManus<br />

1981).<br />

An NPV was recorded from gypsy moth larvae in southeastern Ontario during studies<br />

<strong>of</strong> gypsy moth parasites carried out in 1974. 1975. 1979, and 1980 (Griffiths 1977.<br />

Griffiths & Wallace in press). It was also recorded in southwestern Quebec by Cardinal<br />

& Smirn<strong>of</strong>f (1973). Smirn<strong>of</strong>f has pointed out that for the last 3 years NPV has lowered<br />

gypsy moth numbers in Quebec (W.A. Smirn<strong>of</strong>f 1981 personal communication).


306 K. J. Griffiths and F. W. Qucdnau<br />

Table 79<br />

Releases and Recoveries<br />

Anastatus dis paris<br />

Rusdaka (Hymenoptera:<br />

Eupelmldae)<br />

Parasitoids recovered from the gypsy moth, Lymantria dispar (L.), in <strong>Canada</strong> 1974-80<br />

Madrid & Quednau Griffiths<br />

Griffiths Stewan 1978-80 1979-80<br />

1977 1980 (unpublishcd) (unpublishcd)<br />

Diptera<br />

Sarcophagidae<br />

Sarcophaga aldrichi (Park.) xx xx<br />

Tachinidae<br />

Blepharipa pralensis (Mg.)' xx xx xx<br />

Compsilura concinnala (Mg.)' xx xx xx xx<br />

Exorisla larvarum (L.)' xx xx<br />

Exorisla mella (Wlk.) xx<br />

Paraseligena silvestris (R-D)' xx xx xx xx<br />

Hymenoptera<br />

Braconidae<br />

COlesia melanosceills (Ratz.)· xx xx xx xx<br />

Chalcididae<br />

Brachymeria intermedia (Nees)' xx xx<br />

Encyrtidae<br />

Ooencyrtus kllvanae (How.)" xx<br />

Eupelmidae<br />

AnaslalllS disparis Ruschka" xx<br />

Ichneumonidae<br />

ExochllS sp. xx<br />

Iloplectis conquislilor (Say) xx xx<br />

Phobocampe disparis (Vier.)· xx xx xx<br />

Pimpla pedalis (Cress.) xx xx xx xx<br />

77reronia llIilfattJae fulvescens (Cress.) xx<br />

Theronia hilaris (Say) xx<br />

Scelionidae<br />

TelonomllS sp. xx<br />

• Exotic parasitoids established in the United States .<br />

This parasitoid is a European species, successfully established in the United States by<br />

Howard & Fiske (1911)_ Natural dispersal is slow because female adults cannot fly and the<br />

species was subsequently distributed by further releases throughout the gypsy moth's<br />

American range (Dowden 1962).<br />

Adults <strong>of</strong> this species emerge from the previous year's gypsy moth eggs at about the<br />

time female moths are ovipositing. Female parasitoids attack the eggs and the parasitic<br />

larvae can develop in all embryonic stages <strong>of</strong> the gypsy moth. However, the proportion<br />

<strong>of</strong> gypsy moth females produced and the fecundity and longevity <strong>of</strong> the ovipositing<br />

female is reduced only if older eggs are attacked. There is usually only one generation<br />

per year and the parasitoid overwinters as a mature larva within the host egg. Preliminary<br />

work on the overwintering ability <strong>of</strong> mature larvae <strong>of</strong> the parasitoid indicated that A.<br />

dis paris apparently could survive over the same geographic area in <strong>Canada</strong> as the<br />

gypsy moth (Sullivan el 01. 1977). Because parasitism up to 28% had been recorded by<br />

Burgess & Crossman (1929), introduction <strong>of</strong> this species into <strong>Canada</strong> was undertaken


Table 80<br />

Table 81<br />

Ooenc)'rtus kuvsnse<br />

(How.) (Hymenoptera:<br />

Encyrtidae)<br />

Lymantria clispur (L.). 307<br />

following preliminary field cage and insectary studies (Table 80). The first releases were<br />

made near St-Mathias, Rouville County, Quebec. in 1979. with further releases near<br />

Acton Vale and Drummondville, Bagot and Drummond Counties respectively, in 1980<br />

(Table 81). Establishment <strong>of</strong> A. disparis occurred in all areas where releases were<br />

made. The initial occurrence <strong>of</strong> parasitism was 20-30% when batches <strong>of</strong> 25 female<br />

parasitoids were implanted in small ventilated cages, but only 2% when open releases <strong>of</strong><br />

batches <strong>of</strong> 200-400 females were made. Successful hibernation <strong>of</strong> A. disparis at St­<br />

Mathias was observed in 1980. A third release <strong>of</strong>this parasitoid was made in August 1980<br />

near Kilburnie, Ontario, but there have been no recoveries from this release as yet<br />

(Table 81).<br />

Laboratory and field cage studies <strong>of</strong> parasitoids against the gypsy moth, Lymanrria<br />

dis par (L.) in Quebec<br />

Species Year<br />

AnUSlalUS disparis Ruschka 1979<br />

• Field cage studies.<br />

•• Laboratory and insectary studies.<br />

1980<br />

Origin<br />

Hungary<br />

Austria<br />

Romania<br />

Hungary<br />

Austria<br />

Romania<br />

Hungary<br />

Number<br />

845*<br />

2505**<br />

234"<br />

Open releases and recoveries <strong>of</strong> parasitoids against the gypsy moth, Lymanlria dispar (L.)<br />

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

Species and province Year Origin Number recovery<br />

Anuslalus disparis Ruschka<br />

Quebec 1980 USA<br />

Hungary<br />

5600 1980<br />

Ontario<br />

Ooencyrtus kuvanae (How.)<br />

1980 Hungary 1020<br />

Ontario 1976 USA 25000 1978<br />

This Japanese egg parasitoid was introduced into New England in 1909. Its establishment<br />

was immediate and its natural dispersal was augmented by additional large releases<br />

throughout the infested area (Burgess & Crossman 1929).<br />

In the United States O. kuvanae can have three generations per year on gypsy moth.<br />

The first generation cycle takes approximately 6 weeks but the subsequent ones only 3<br />

weeks. Early studies showed that attack by this species was weak. Britton (1935)<br />

recorded that in southeastern Massachusetts and in Rhode Island egg parasitism was not<br />

more than 10% per year. However, higher attack levels have been recorded in more


308 K. J. Griffiths and F. W. Qucdnau<br />

Borrelinsyirus<br />

reprlmens Holmes<br />

Bad11us tburingiemis<br />

Berliner (B.t.)<br />

Evaluation <strong>of</strong> Control Attempts<br />

recent work. Dowden (1962) found attacks <strong>of</strong> 40-45% in Massachusetts and Connecticut.<br />

Weseloh (1972) gives a total summer attack <strong>of</strong> 29-42% in Connecticut. Thus its<br />

introduction into <strong>Canada</strong> was justified and preliminary investigations <strong>of</strong> the ability <strong>of</strong> the<br />

overwintering mated female adult to withstand low temperatures were undertaken<br />

(Griffiths & Sullivan 1978). It was found that adults were unable to survive continuous<br />

exposure to O°C for 30 days, and it was concluded that this species would be incapable <strong>of</strong><br />

surviving in the current range <strong>of</strong> the gypsy moth in <strong>Canada</strong>. To test this hypothesis,<br />

approximately 25 000 adults supplied by the New Jersey Department <strong>of</strong> Agriculture<br />

were released on Wolfe Island, Ontario, on 24 September 1976 (Table 81). Gypsy moth<br />

egg masses have been collected in the release area in late August or early September<br />

each year since the release. There were no recoveries <strong>of</strong> O. kuvanae until 1978, when 28<br />

adults were recovered from two egg masses. A single adult was obtained from the<br />

release area in 1980. In addition, 45 adults <strong>of</strong> O. kuvanae were recovered from two egg<br />

masses collected near Glen Norman, Glengarry County, Ontario, in September 1980.<br />

Because this recovery site is approximately 180 km from the Wolfe Island release site, it<br />

is unlikely that there is any relation between the two; the parasitoids must have dispersed<br />

there naturally.<br />

Two experimental techniques for the application <strong>of</strong> NPV were tested by Cardinal &<br />

Smirn<strong>of</strong>f (1973) in Missisquoi and St-Jean Counties, Quebec, in 1972. They handapplied<br />

an aqueous solution <strong>of</strong> NPV at 40 x 10' polyhedral inclusion bodies (PI B) per<br />

millilitre to egg masses in early May and recorded nearly 100% mortality in the larvae<br />

emerging from them. They also sprayed foliage by mist blower with an aqueous solution<br />

at 4 x 10' PIBlml when larvae were in the second or third instar, resulting in higher<br />

mortality than was seen in larvae on unsprayed foliage.<br />

Jobin & Caron (1982) made aerial applications <strong>of</strong> B.I., using the commercial product<br />

Thuricide® 32B (Sandoz Inc.) in Chambly and Rouville Counties, Quebec, in 1979.<br />

There were two applications in each area at dosages <strong>of</strong> 19.6 and 39.52 x W"I.U.fha. All<br />

applications included a sticker and chitinase and the B.I. was applied at a rate <strong>of</strong> 9.36<br />

Vha. Spraying was carried out in late May when 60-70% <strong>of</strong> larvae were in the second<br />

instar. Mortality <strong>of</strong> larvae in the treated areas 20 days after treatment was 100%,<br />

compared to a decrease <strong>of</strong> 70-76% in control areas over the same time period.<br />

The establishment <strong>of</strong> exotic insect parasites on the gypsy moth in <strong>Canada</strong> is proceeding<br />

well, largely through natural dispersal. We now know that 8 <strong>of</strong> the 13 established exotic<br />

species are attacking gypsy moth in the infested zone in <strong>Canada</strong> and we are confident<br />

that a 9th species, Anaslalus disparis Ruschka, will become established. Two <strong>of</strong> the<br />

remaining four species are already present in <strong>Canada</strong> on other hosts. They are Apanleles<br />

klcteicolor Viereck, essentially a parasitoid <strong>of</strong> the browntail moth, which was established<br />

against that species in the maritimes and Quebec in 1915 (Tothill 1916); and M. versicolor,<br />

introduced against browntail moth and satin moth in <strong>Canada</strong> (TothillI916) and recovered<br />

in Ontario from Rlleumaplera lIaslala L. by Forest Insect and Disease Survey staff. M.<br />

versicolor is infrequently recovered from the gypsy moth in the United States. The<br />

other two exotic species are Eupleromalus hemiplerus (Walker) and Monodonlomerus<br />

aereus Walker. The former species is a common parasitoid <strong>of</strong> the satin moth in the<br />

United States but is rarely obtained from the gypsy moth there (Burgess & Crossman


Recommendations<br />

Literature Cited<br />

Lymantria dispar (L.). 3()9<br />

1929, Proper 1931). nle latter species is a secondary parasitoid as frequently as it is a<br />

primary one (Howard & Fiske 1911).<br />

It is too soon to state that the release <strong>of</strong> egg parasitoids in <strong>Canada</strong> has resulted in the<br />

permanent establishment <strong>of</strong> these insects. We do know that seven other species <strong>of</strong><br />

exotic parasitoids previously established in the United States are now attacking gypsy<br />

moth in <strong>Canada</strong>, but, in the absence <strong>of</strong> careful population studies <strong>of</strong> the gypsy moth we<br />

cannot say what impact any <strong>of</strong> them is having.<br />

A combination <strong>of</strong> applications <strong>of</strong> carbaryl on some areas and <strong>of</strong> "insecticide soap" on<br />

other areas in 1979 in Vancouver apparently resulted in elimination <strong>of</strong> the population<br />

there. It should be pointed out, however, that very few <strong>of</strong> the isolated infestations<br />

recorded in the history <strong>of</strong> the gypsy moth in the United States have ever been completely<br />

eradicated. Spraying with either an NPV or B.I. has been effective in small scale trials<br />

but the cost <strong>of</strong> application is many times higher than the cost <strong>of</strong> applying chemicals.<br />

It is evident that, aside from continuing to release and sample for the two egg parasitoids, O.<br />

kuvanae and A. dis paris , there is little more to be done in the introduction <strong>of</strong> biological<br />

agents because there are no more known suitable candidates. The one successfully<br />

established exotic predator, Calosoma sycophanla L., was introduced unsuccessfully<br />

into <strong>Canada</strong> at a number <strong>of</strong> locations (McGugan & Coppe11962) and it is questionable<br />

whether further work on this species is justified.<br />

Cardinal & Smim<strong>of</strong>f (1973) were able to produce gypsy moth NPV by rearing<br />

contaminated larvae in cages in the field. Smim<strong>of</strong>f (personal communication) recommends<br />

that the NPV so produced be used to inoculate incipient outbreaks by application on<br />

either egg masses or foliage to prevent buildup <strong>of</strong> gypsy moth to outbreak levels.<br />

In view <strong>of</strong> the above and <strong>of</strong> the relatively innocuous levels <strong>of</strong> gypsy moth populations<br />

in <strong>Canada</strong> in 1980, we feel there is no urgent need for further classical biological control<br />

work on this species. However, we strongly recommend that careful surveillance <strong>of</strong> the<br />

dispersal <strong>of</strong> the gypsy moth be continued by staff <strong>of</strong> the Plant Protection Division <strong>of</strong><br />

Agriculture <strong>Canada</strong>. Continued monitoring in the Vancouver area is especially important.<br />

Brillon. W.E. (1935) The gypsy moth. Connecticut Agricultural Experiment SUllion Bulletin 375,623-647.<br />

Brown, G.S. (1968) The gypsy moth, Porthetria dispar L., a threat to Ontario honicullure and forestry. Proceedings <strong>of</strong> the <strong>Entomological</strong><br />

<strong>Society</strong> <strong>of</strong> Ontario 98,12-15.<br />

Burgess, A. F. (1914) The gypsy moth and the brown· tail moth. with suggestions for their control. US Department <strong>of</strong> Agriculture Farm Bulletin<br />

564,24 pp.<br />

Burgess, A.F.; Crossman. S.S. (1929) Imponed insect enemies <strong>of</strong> the gypsy moth and the brown-tail moth. US Departnunt <strong>of</strong> Agriculture<br />

Technical Bulletin 86, 147 pp.<br />

Campbell, R.W.; Podgwaite. J.D. (1971) The disease complex <strong>of</strong> the gypsy moth. I. Major components. Journal <strong>of</strong> In"ertebrate Pathology<br />

18(1).101-107.<br />

Cardinal. J .A.; Smim<strong>of</strong>f. W.A. (1973) Introduction ex¢rimentale de la polyCdrie nucleairc de Portherria di.fpar L. (Upidoptcres: Lymantriidae)<br />

en foret. Phytoprotection 54(1).48-50.<br />

Doane. C.C. (1970) Primary pathogens and their role in the development <strong>of</strong> an epizootic in the gypsy moth. Journal <strong>of</strong> Invertehrate Pathology<br />

15(1),21-23.<br />

Doane. c.c. (1971) Field application <strong>of</strong> a Streptococcus causing brachyosis in larvae <strong>of</strong> Porthetria dis par. Journal <strong>of</strong> In"enehrate Pathology<br />

17(3),303-307.<br />

Doane. c.c.; McManus. M.L. (Eds.) (1981) The gypsy moth: research toward integrated pest management. US Department <strong>of</strong> Agriculture<br />

Science Education Agency APHIS Technical Bulletin 1584,757 pp.<br />

Dowden. P.B. (1962) Parasites and predators <strong>of</strong> forest insects liberated in the United States through 1960. US Department <strong>of</strong> Agriculture Forest<br />

Se",ice Handbook 226. 70 pp.<br />

Griffiths. K.J. (1976) The parasites and predators <strong>of</strong> the gypsy moth: a review <strong>of</strong> the world literature with special application to <strong>Canada</strong>.<br />

Canadian Forestry Se",ice Sault Ste. Marie. Ontario Report O-X·243. 92 pp.


Pest Status<br />

Background<br />

Chapter 55<br />

Malacosoma disstria Hubner, Forest Tent<br />

Caterpillar (Lepidoptera: Lasiocampidae)<br />

W.G.H.IVES<br />

The forest tent caterpillar, Malacosoma disstria HUbner, has had a long history <strong>of</strong><br />

periodic outbreaks in <strong>Canada</strong> (Baird 1917, Hodson 1941, Hildahl & Reeks 1960, Sippell<br />

1962, Witter et al. 1975). The insect attacks a wide variety <strong>of</strong> tree species (Prentice<br />

1963), although the preferred host in most areas is trembling aspen, Populus tremu/oides<br />

Michx. Host trees are <strong>of</strong>ten completely stripped <strong>of</strong> foliage during outbreaks, but because<br />

the outbreaks usually last for only 3-6 years at anyone location (Sippell 1962, Witter et<br />

al. 1975), and even severely defoliated trees usually refoliate within a few weeks, there is<br />

usually very little tree mortality directly attributable to defoliation (Kulman 1971). However,<br />

severe mortality has occasionally been reported (Hodson 1941, G.N. Sti111980<br />

personal communication). Observations by the author suggest that such mortality is<br />

attributable primarily to the effects <strong>of</strong> severe drought fonowing several years <strong>of</strong> moderate to<br />

severe defoliation.<br />

Between 1969 and 1980 records compiled by the Forest Insect and Disease Survey<br />

(Anon. 1970-79) show that outbreaks <strong>of</strong> the forest tent caterpillar have occurred in all<br />

provinces except Newfoundland (Fig. 16). In New Brunswick, about 9 000 ha <strong>of</strong> trembling<br />

aspen were severely defoliated in 1969, and the infestation was expected to enlarge in<br />

1970. However, a combination <strong>of</strong> unseasonably wann weather in early May 1970, fonowed by<br />

several frosts, apparently resulted in the starvation <strong>of</strong> many young larvae. The area <strong>of</strong><br />

infestation therefore decreased, and the expected outbreak did Dot materialize. In Nova<br />

Scotia, about 50 km 2 <strong>of</strong> aspen were defoliated in 1972, and in Prince Edward Island about<br />

28 000 ha in 1974. In Quebec, several small areas totalling less than 100 km 2 were<br />

defoliated between 1973 and 1976. In Ontario, remnants <strong>of</strong> earlier outbreaks persisted<br />

into the survey period and new outbreaks began developing soon afterwards. A total<br />

<strong>of</strong> over 100 000 km 2 east <strong>of</strong> the Lakehead was defoliated between 1975 and 1977, while<br />

the outbreak in northwestern Ontario alone covered about 160 000 km 2 in 1978. In<br />

Manitoba, an incipient outbreak was detected in 1971 and by 1977 most <strong>of</strong> the southern<br />

part <strong>of</strong> the province was affected. The outbreak collapsed in 1978, except for an area<br />

west <strong>of</strong> Lake Winnipegosis where it collapsed in 1979. In Saskatchewan, the outbreak<br />

was slower in developing and did not reach its peak until 1980, when about 128 000 km 2<br />

were subjected to moderate or severe defoliation. In Alberta, an outbreak in the Wabamun<br />

Lake area has persisted throughout the survey period, but the greatest damage occurred<br />

in 1980, when about 75000 km 2 were defoliated. In British Columbia, over 70 000 ha<br />

were defoliated in the central part <strong>of</strong> the province in 1973, but the outbreak collapsed in<br />

1974.<br />

Populations <strong>of</strong> forest tent caterpillar are influenced by a variety <strong>of</strong> environmental factors<br />

including weather, competition, predation, parasitism, and diseases. Ives (1973) believed that<br />

unusually favourable spring weather 2-4 years before any noticeable defoliation was<br />

responsible for triggering outbreaks, and a number <strong>of</strong> workers report that unfavourable<br />

spring weather appears to be a major factor in population collapse (Tothilll918, Sweetman<br />

311


Malacosoma disstria J-Hibner. 3/3<br />

1940, Blais et al. 1955, Witter et al. 1972, Hodson 1977). Unusually cold winter weather<br />

is also sometimes responsible for population decline (Witter et al. 1975). Hodson (1941,<br />

1977) found that starvation, <strong>of</strong>ten coupled with high pupal parasitism by Sarcophaga<br />

aldrichi Parker, was a major contributor to population collapse. A large variety <strong>of</strong> insect<br />

parasites attacks the forest tent caterpillar (Witter & Kulman 1972), but the most prevalent is<br />

usually S. aldrichi, which <strong>of</strong>ten attacks over 90% <strong>of</strong> the pupae in the declining phases <strong>of</strong><br />

an outbreak (Hodson 1941, 1977). Predation by various birds (McAtee 1926, Hodson<br />

1941, Witter & Kulman 1972) and invertebrate predators such as ants (Tot hill 1918,<br />

Green & Sullivan 1950, Ayre & Hitchon 1968) has been observed on a number <strong>of</strong> occasions,<br />

but it is usually unimportant.<br />

Bacteria, fungi, microsporidia, and viruses all cause diseases in forest tent caterpillar<br />

larvae and pupae (Bird 1971). The bacteria isolated from M. disstria larvae include<br />

Bacillus cereus Frankland & Frankland, Clostridium brevifaciens Butcher, a Pseudomonas<br />

species, and Serratia marcescens Bizio, but all were rare (Smim<strong>of</strong>f 1968). The<br />

forest tent caterpillar is also susceptible to Bacillus thuringiensis Berliner (B.t.), although<br />

this organism is not usually found under natural conditions. Several fungi have been<br />

isolated from forest tent caterpillar larvae and pupae, but according to Stairs (1972) only<br />

Beauveria bassiana (Balsamo) Vuillemin and Entomophthora sp. were <strong>of</strong> any importance,<br />

and then only under humid conditions. A microsporidium, Nosema disstriae Thompson,<br />

has killed up to 90% <strong>of</strong> early-instar larvae in some areas (Smirn<strong>of</strong>f 1968). It also causes a<br />

reduction in the size <strong>of</strong> the insects and may affect their vigour, so that the full significance <strong>of</strong><br />

infection is hard to assess (Bird 1971).<br />

Infection with a nuclear polyhedrosis virus (NPV) is <strong>of</strong>ten thought to be one <strong>of</strong> the<br />

most important factors in terminating forest tent caterpillar outbreaks. For example,<br />

Stairs (1966) stated that "Epizootics <strong>of</strong> nuclear polyhedrosis virus disease are known to<br />

be associated with the rapid decline <strong>of</strong> tent caterpillar popUlations, Malacosoma spp."<br />

Although this statement may be true for the tent-forming species, the evidence is<br />

equivocal for the forest tent caterpillar. Of the references cited by Stairs (1966), the<br />

following three deal with M. disstria. Bergold (1951) obtained high mortality among<br />

larvae in cages when they were sprayed with the virus under experimental conditions.<br />

Sippell (1952) demonstrated that a large percentage <strong>of</strong> egg bands collected in late fall<br />

were infected with virus. Neither experiment attempted to show that the virus had an<br />

effect on population trends. The third paper (Chapman & Glaser 1915) states: "Though<br />

wilt disease was also reported to have occurred in many places in the forest tent<br />

caterpillar, neither <strong>of</strong> us have seen more than a few typical cases from the field. Many<br />

caterpillars were sent in by field men but only a few <strong>of</strong> them proved to have typical wilt."<br />

Smim<strong>of</strong>f (1968) conducted a long-term study <strong>of</strong> M. disstria diseases in Quebec, and<br />

found NPV to be unimportant. Similarly, population studies by Hodson (1941) showed<br />

that the virus was responsible for less than 1 % mortality in Itasca State Park in 1936, and<br />

Witter el al. (1972), on the basis <strong>of</strong> intensive sampling, considered the virus to be<br />

unimportant in northern Minnesota during 1967-69.<br />

Although NPV appears to be unimportant as a control agent under natural conditions,<br />

it is possible that artificial dissemination <strong>of</strong> the virus into field populations might<br />

increase the amount <strong>of</strong> virus infection, thus increasing its effectiveness. Stairs (1964)<br />

used a small backpack mist blower to apply various dosages <strong>of</strong> virus to different larval<br />

instars <strong>of</strong> the forest tent caterpillar in 1963. A concentration <strong>of</strong> 10' polyhedral inclusion<br />

bodies (PIB) per millilitre applied to second- and third-instar larvae caused 92% mortality,<br />

most <strong>of</strong> the larvae dying within 10 days <strong>of</strong> the application. Stairs (1965) found virusinfected<br />

larvae in 1964 in both <strong>of</strong> the areas that had been sprayed in 1963, but not in an<br />

unsprayed area. The virus also occurred at considerable distances from the sprayed<br />

areas, and Stairs believed that he had been successful in artificially initiating an epizootic.


314 W. O. H. Ives<br />

Releases and Recoveries<br />

Bacillus tburingJensis<br />

Berliner<br />

Biological control attempts against the forest tent caterpillar during the past decade have<br />

involved the bacterium B. Ihuringiensis, the microsporidium N. disslriae, and an NPV.<br />

Because some formulations <strong>of</strong> B. Ihuringiensis have been registered for control <strong>of</strong> the<br />

forest tent caterpillar, and have therefore been used commercially it is difficult to provide<br />

an accurate record <strong>of</strong> either the amounts used or the effectiveness <strong>of</strong> the applications. The<br />

following is a summary <strong>of</strong> the more readily available information.<br />

The most comprehensive trials <strong>of</strong> B.I. against the forest tent caterpillar were conducted<br />

in Ontario in 1975,1977, and 1978 (G.M. Howse 1981 personal communication). In 1975,<br />

small-scale aerial spray trials with Thuricide® 168 (Sandoz Inc.) on trembling aspen<br />

stands in southern Ontario showed that the preparation had promise. One application at a<br />

rate <strong>of</strong> 10 x 10 9 1. U .lha reduced larval populations by 71 %; two applications at the same<br />

rate caused 92% mortality. The treatment was applied too late in the season to save any<br />

foliage. In 1977, evaluations were made <strong>of</strong> the effectiveness <strong>of</strong> two applications <strong>of</strong><br />

Thuricide® 16B (10 x 10' I. U .lha at each application) in controlling forest tent caterpillar<br />

populations over a total <strong>of</strong> about 850 ha <strong>of</strong> trembling aspen and sugar maple, Acer<br />

saccharum Marsh., in three provincial parks in southern Ontario. Resultant population<br />

reduction on trembling aspen ranged from 30 to 79%, and defoliation in the treated areas<br />

ranged from 5 to 46% compared to 82 to 100% in untreated areas. Similar results were<br />

obtained for infestations on sugar maple, where population reductions due to treatment<br />

ranged from 38 to 69%, and defoliation in sprayed areas was about 15%, compared to<br />

100% in an unsprayed area. A further 940 ha were sprayed with Thuricide® 16B in 1977,<br />

but no assessment <strong>of</strong> its effectiveness was made. In 1978, either one or two applications <strong>of</strong><br />

Thuricide® 168, mostly at the rate <strong>of</strong>7.5 x 10' I. U .lha for each application, were made on<br />

a total <strong>of</strong> about 515 ha <strong>of</strong> forest tent caterpillar infestations on trembling aspen, sugar<br />

maple, and red oak, Quercus rubra L., in southern Ontario. Insect populations were<br />

lower than in 1977, so the comparison <strong>of</strong> defoliation was not a practical way <strong>of</strong> showing<br />

effectiveness. However, population reduction due to treatment ranged from 79 to 100%<br />

on trembling aspen, from 71 to 100% on red oak, and was 100% in both treated<br />

infestations on sugar maple. In addition, 250 ha <strong>of</strong> sugar maple were sprayed with one<br />

application <strong>of</strong> Dipel® WP (Abbott Ltd.) (6.0 x 10' I.U.lha) in late May. Population<br />

reduction due to treatments was 98%.<br />

Data for western <strong>Canada</strong> are more fragmentary. In Alberta, a hydraulic sprayer was<br />

used in 1974 to treat two 0.2-ha plots <strong>of</strong> trembling aspen with Dipel® WP (9 x 10'<br />

I.U.lha) and a further two 0.2-ha plots with Thuricide® HPC (Sandoz Inc.) (5 x ur<br />

I.U.lha) (Drouin & Kusch 1975). Control was estimated at 75 and 70% respectively. In<br />

1976, two O.4-ha plots <strong>of</strong> trembling aspen were treated with Dipel® WP (9 x 10'<br />

I.U./ha). Control was estimated at 75 and 85%, based on the reduction in amount <strong>of</strong><br />

defoliation (Drouin & Kusch 1977). In addition, several applications were made by<br />

private operators under contract to various agencies.<br />

In Saskatchewan, Dipel® WP was used to control forest tent caterpillar outbreaks in<br />

Prince Albert and Saskatoon. In Prince Albert, about 325 ha within the city limits were<br />

treated in 1980 when the larvae were in the third and fourth instars, with good results (P.<br />

Kabat<strong>of</strong>f 1981 personal communication). The chemical insecticide malathion had been<br />

used earlier, when the larvae were in the first and second instars, because it was felt that<br />

there was insufficient foliage at that time for Dipel® WP to be effective. In Saskatoon,<br />

about 800 ha <strong>of</strong> city parks and about 40 000 avenue trees were treated with Dipel® WP in<br />

1980 (D. Scott 1981 personal communication). About 340 kg <strong>of</strong> Dipel® WP were used,<br />

and results were generally satisfactory.


Malam.mma disstria Hiibner, 319<br />

Stairs, O.R. (1964) Dissemination <strong>of</strong> nuclear polyhedrosis virus agaimt the forest tent caterpi11ar, MoJorosomo disstrio (Hiibner) (Lepidoptera:<br />

Lasiocampidae). Canadian EntomologisI96,1017-1020.<br />

Stairs, O.R. (196S) Artificial initiation <strong>of</strong> virus epizootics in forest tent caterpillar populations. Canadian Entomologist 97,10S9-1062.<br />

Stairs, O.R. (1966) Transmission <strong>of</strong> virus in tent caterpillar populations. Canadian Entomologist 98,1100-1104.<br />

Stairs, O.R. (1972) Pathogenic microorganisms in the regulation <strong>of</strong> forest insect populations. AnnUQI Review <strong>of</strong> Entomology 17,3SS-372.<br />

Sweetman, H. L. (1940) The value <strong>of</strong> the hand control for the tent caterpillars, Malacosoma americana Fabr. and Malacosoma di.s.stria Hbn.<br />

(Lepidoptera: Lasiocampidae). Canadian Enlomologis/72.24S-250.<br />

Tothill. J.D. (1918) The meaning <strong>of</strong> natural contro\. Proceedings <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> 4,10-14.<br />

Wilson. G.O. (1977) Effects <strong>of</strong> the microsporidia Nosema disstriae and Pleistophora schubergi on the survival <strong>of</strong> the forest tent caterpillar,<br />

Malacosoma difstria (Lepidoptera: Lasiocampidae). Canadian Entomologisl 109,1021-1022.<br />

Wilson, 0.0. (1979) Effects <strong>of</strong> Nosema disstriae (Microsporidia) on the forest tent caterpillar, Malacosomo disstrio (Lepidoptera: Lasiocampidae).<br />

Proceedings <strong>of</strong> Ihe <strong>Entomological</strong> Socitty <strong>of</strong> Ontario 110.97-99.<br />

Wilson. G.O.; Kaupp. W.J. (1977) Application <strong>of</strong> Nosema disstriat and Pleistophora schubergi (Microsporida) against the forest tent<br />

caterpillar in Ontario. 1977. Canadian Forestry Service, Sault 511'. Marie, Information Report FPM-X-4.<br />

Willer. I.A.: Kulman. H.M. (1972) A rcview <strong>of</strong> the parasites and predators <strong>of</strong> tent caterpillars (Malacosoma spp.) in North America.<br />

University <strong>of</strong> Minnesola Agricultural Experiment Stalion Technical Bulletin 289.<br />

Witter, J .A.; Kulman, H .M.; Hodson, A.C. (1972) Life tables for the forest tent caterpillar. Annals <strong>of</strong> the EntomologiCilI <strong>Society</strong> <strong>of</strong> America<br />

6S.25-3t.<br />

Willer, J .A.; Mattson, W.J .• Kulman. H .M. (197S) Numerical analysis <strong>of</strong> a forest tent caterpillar (Lepidoptera: Lasiocampidae) outbreak in<br />

northern Minnesota. Canadian Entomologist 107,837-854.


Blank Page<br />

320


Pest Status<br />

Background<br />

Field Trial<br />

Chapter 56 321<br />

Neodiprion abietis (Harris), Balsam Fir<br />

Sawfly (Hymenoptera: Diprionidae)<br />

J.e. CUNNINGHAM<br />

The balsam fir sawfly, Neodiprion abietis (Harr.) complex, occurs from New England to<br />

the Great Lakes and south to Missouri in the USA, from coast to coast in southern<br />

<strong>Canada</strong> (Baker 1972), and also in Newfoundland. In <strong>Canada</strong>, periodic outbreaks, usually <strong>of</strong><br />

short duration, are common on balsam fir, Abies balsamea (L.) Mill., and spruce, Picea<br />

spp., from Saskatchewan eastwards. Tree mortality is rare and is usually reported in<br />

conjunction with other pests such as spruce budworm, Choristoneura fumiferana<br />

(Clem.), eastern blackheaded budworm, Ac/eris variana (Fern.), and balsam woolly<br />

adelgid, Adelges piceae (Ratz.) (Rose & Lindquist 1977).<br />

Balsam fir sawfly overwinters as eggs in diapause and spring hatching is fairly<br />

uniform. Larvae are gregarious and feed on old needles, but eat only parts <strong>of</strong>them, thus<br />

damaging many more needles than they consume. In the later instars, the colonies<br />

disperse and the larvae become solitary. There are five larval instars, but males do not<br />

feed during their last instar (D.R. Wallace & c.R. Sullivan 1980 personal communication).<br />

In 1980, traces <strong>of</strong> defoliation caused by this insect were found in western and central<br />

Newfoundland and light to moderate defoliation was observed in the northern, eastern,<br />

and Algonquin regions <strong>of</strong> Ontario (Sterner & Davidson 1981).<br />

A nuclear polyhedrosis virus (NPV) was first reported in balsam fir sawfly by Steinhaus<br />

(1949). This disease has subsequently been reported in balsam fir sawfly populations in<br />

several Canadian provinces including Saskatchewan (Brown 1951), Alberta and Manitoba<br />

(Cumming 1954), and Quebec (Martineau & Lavallee 1971).<br />

Ground spray trials on single trees were conducted in Ontario in 1972 to test NPV on<br />

balsam fir sawfly (Ol<strong>of</strong>sson 1973). Aqueous suspensions <strong>of</strong> lOT, 1()6, lOS, 10', and to)<br />

polyhedral inclusion bodies (PIB) per millilitre plus 2.5% IMC 90-001 (International<br />

Minerals Corporation) UV protectant were tested at two application times, 8 days<br />

apart, on first- and third-instar larvae. A backpack mist blower was used and a 360-ml<br />

suspension was applied to each 2.5-m tall balsam fir tree, five trees being treated with<br />

each dosage on each date. There were 5-10 sawfly colonies per tree.<br />

The percentage <strong>of</strong> infected larvae following the applications is shown in Table 83.<br />

Infection is directly related to mortality with this NPV.


322 J. C. Cunningham<br />

Table 83<br />

Recommendations<br />

Literature Cited<br />

Infection and defoliation following application <strong>of</strong> NPV on first- and third-instar balsam fir<br />

sawfly, Neodiprion abietis (Harr.), larvae in a ground spray trial (from Ol<strong>of</strong>sson 1973)<br />

Percentage virus infection Defoliation·<br />

Dosage I instar III instar I instar III instar<br />

PIBlml treated treated treated treated<br />

10' Q<br />

100 100 1 3<br />

10'<br />

HY<br />

100<br />

96<br />

88<br />

60<br />

2<br />

3<br />

4<br />

5<br />

10'<br />

10}<br />

72<br />

60<br />

24<br />

4<br />

4<br />

4<br />

5<br />

5<br />

Check 0 0 5 5<br />

• Defoliation was rated on a scale <strong>of</strong> 1 to 5 where 1 was very light and 5 severe.<br />

The limited ground spray trials conducted in 1972 demonstrated that this NPV is a<br />

potentially useful biological control agent for balsam fir sawfly. However, as this<br />

species feeds mainly near the tops and mid-crowns <strong>of</strong> mature trees, an aerial spray<br />

application would be required for complete coverage. Further testing has not been<br />

conducted and aerial spray trials are required before definitive conclusions can be made<br />

and recommendations formulated.<br />

Ol<strong>of</strong>sson (1973) showed that timing <strong>of</strong> the application is critical and he recommended<br />

applying the NPVon late first-instar larvae. By the time they develop to the third instar,<br />

feeding damage is severe and the larvae require a much higher dosage <strong>of</strong> virus. For<br />

ground spray applications, Ol<strong>of</strong>sson recommended a dosage <strong>of</strong> 1W PIB/ml and he<br />

calculated that one fully grown, NPV-infected larva produced 1 I <strong>of</strong> spray <strong>of</strong> this<br />

concentration. However, this dosage would have to be increased for an aerial application. If<br />

the economic importance <strong>of</strong> balsam fir sawfly justifies the development <strong>of</strong> non-chemical<br />

pesticide control strategies, this NPV merits further research.<br />

Baker, W.J. (1972) Eastern forest insects. US Depanment <strong>of</strong> Agriculture Miscellaneous Publication 1175. US Government Printing OCfice,<br />

642 pp.<br />

Brown, C.E. (1951) Forest insect survey nOles. Canadian Depanment <strong>of</strong> Agriculture Bi·monthly Progress Report 7(2).2.<br />

Cumming. M. (1954) Diseases <strong>of</strong> insects. Canadian Depanment <strong>of</strong> Agriculture Bi·monthly Progress Report 10(3),3.<br />

Martineau, R.; Lavallc!e, A. (1971) Quc!bec region. Canadian Forestry Service, Annual Report <strong>of</strong> the Forest Insect and Disease Survey,<br />

pp.34-53.<br />

OloCsson, E. (1973) Evaluation <strong>of</strong> a nuclear polyhedrosis virus as an agent for the control <strong>of</strong> balsam fir sawfly, Neodiprion abietis (Harr.).<br />

Canadian Forestry Service, Sault Ste. Marie, Information ReportIP·X·2, 30 pp.<br />

Rose, A.H.; Lindquist, O.H. (19n) Insects <strong>of</strong> eastern spruces, fir and hemlock. Canadian Forestry Service, Forestry TechniCDI Report<br />

23, 159 pp.<br />

Steinhaus, E.A. (1949) Principles <strong>of</strong> insect pathology. New York; McGraw· Hill, 757 pp.<br />

Sterner, T.E.; Davidson, A.G. (comp.) (1981) Forest insect and disease conditions in <strong>Canada</strong> 1980. Canadian Forestry Service, Forest<br />

Insect and Disease Survey, 43 pp.


Pest Status<br />

Background<br />

Field Trials<br />

Chapter 57<br />

Neodiprion lecontei (Fitch), Redheaded<br />

Pine Sawfly (Hymenoptera: Diprionidae)<br />

J.C. CUNNINGHAM and P. DE GROOT<br />

The redheaded pine sawfly, Neodiprion lecontei (Fitch), is a serious defoliator <strong>of</strong> young<br />

hard pine, Pinus spp., plantations in Ontario, Quebec, and, to a lesser extent, New<br />

Brunswick. It is the most serious pest <strong>of</strong> red pine, Pinus resinosa Ait., in these regions<br />

due to the increase in the number <strong>of</strong> plantations over the last 20 years. Red pine is the<br />

principal species attacked, but jack pine, P. banbiana Lamb., and Scots pine, P.<br />

syfveslris L., are also attacked. Damage can be severe, especially when trees are less<br />

than 2-3 years old; one sawfly colony can totally defoliate and kill several trees<br />

(Benjamin 1955). Heavy defoliation results in reduced height growth, branch mortality,<br />

and defonnity if the leader is killed (Schaffner 1951, Benjamin 1955, MacAloney &<br />

Wilson 1964).<br />

This sawfly has only one generation per year in <strong>Canada</strong>, whereas in the southern parts<br />

<strong>of</strong> its range it may have as many as five (Benjamin 1955). In <strong>Canada</strong>, adult sawflies<br />

emerge, from mid-June to early July, from overwintered cocoons in the duff layer <strong>of</strong> the<br />

soil. The number <strong>of</strong> eggs laid by females varies considerably, but it is not unusual to have<br />

a complement <strong>of</strong> 80-140 eggs. Egg development varies with the ambient temperature<br />

and. as temperatures increase, the incubation period shortens. An incubation period <strong>of</strong><br />

3-4 weeks is typical in southern Ontario and Quebec. The sawfly larvae are gregarious<br />

and feed on foliage <strong>of</strong> all ages, although they show a preference for the previous year's<br />

foliage. Larvae are fully grown in 25-30 days from time <strong>of</strong> hatching, and are about 25<br />

mm long. Populations <strong>of</strong> sawflies are not evenly distributed throughout plantations but<br />

occur in "hot spots", <strong>of</strong>ten along a hardwood perimeter or on knolls.<br />

In the past, most attempts at control have been by applying chemical insecticides from<br />

the ground. In many cases, enough larvae escaped treatment to produce a population the<br />

following year large enough to require further treatment. This "insecticide treadmill"<br />

requires considerable man-power and may have to be continued for several years until<br />

the trees have grown enough to be out <strong>of</strong> danger or until the insect population has been<br />

regulated by natural control factors such as egg parasitoids (Benjamin 1955, Griffiths<br />

1956, Rauf el af. 1979).<br />

In 1950, a nuclear polyhedrosis virus (NPV) disease <strong>of</strong> the redheaded pine sawfly was<br />

found in Ontario (Bird 1961). In a series <strong>of</strong> laboratory and ground spray trials. Bird<br />

demonstrated that this virus was highly infectious to sawfly larvae, that virus can be<br />

introduced into the population and initiate an epizootic, and that it can be transmitted<br />

from one generation to the next (Bird 1961, 1971).<br />

Virus for field trials is produced in vivo. Because there are no known synthetic diets for<br />

sawflies reared in the laboratory, it is necessary to use fresh foliage, which is labour<br />

intensive and expensive. Hence to produce virus, a plantation with a high insect<br />

population, which has already been severely damaged in previous years, is selected.<br />

When larvae reach the fourth instar, trees are sprayed with a mist blower using a<br />

323


324 J. C Cunningham and P. Dc Groot<br />

Table 84<br />

suspension <strong>of</strong> 10" polyhedral inclusion bodies (PIS) per millilitre at an application rate <strong>of</strong><br />

about 20 Ilha. Seven or 8 days after spraying, the first diseased and dead larvae are<br />

found, and colonies <strong>of</strong> moribund larvae are clipped from the trees and taken to the<br />

laboratory where they are picked <strong>of</strong>f the foliage and frozen. Collections are made every<br />

day until all colonies have died or pupated, which can take up to 30 days. The frozen<br />

larvae are lyophilized, ground to a fine powder, and the number <strong>of</strong> PISs per gramme<br />

measured. This material is stored at 2°C until required and is then suspended in water.<br />

Recent experiments have shown that suspension in an emulsifiable oil may be preferable,<br />

but such preparations have not yet been field tested.<br />

Spraying <strong>of</strong> an entomopathogen must be shown to be efficacious and safe before it is<br />

approved for operational use by the Canadian federal government. Extensive safety<br />

testing <strong>of</strong> redheaded pine sawfly NPV has been undertaken at Ontario Veterinary<br />

College, University <strong>of</strong> Guelph, where this virus was found to present no hazard to<br />

mammals (Forsberg, ValIi & Dwyer unpublished), birds (Valli & Oaxton, unpublished),<br />

and fish and Daphnia pulex (Hicks et al. 1981). The impact <strong>of</strong> the virus on non-target<br />

organisms was monitored in conjunction with the 1977 aerial spray trials in Ontario. In<br />

addition, beehives were placed in the treated plantations. No deleterious effects were<br />

noted (Kingsbury et al. 1978).<br />

A petition for the registration <strong>of</strong> this virus will be submitted to Canadian authorities in<br />

March 1981 for approval. If accepted this will be the first \irus registered for regulation<br />

<strong>of</strong> insect populations in <strong>Canada</strong>.<br />

Field trials during the review period started with a smalI ground spray trial with NPV,<br />

conducted in Quebec in 1970. In 1976, an intensive research programme was initiated to<br />

establish an operational method <strong>of</strong> applying this virus to control redheaded pine sawfly.<br />

Aerial spray trials were conducted every year in Ontario from 1976 to 1980, and in<br />

Quebec in 1978 and 1979. Ground applications were also made during this period; the<br />

number <strong>of</strong> plantations and areas treated are summarized in Table 84. Details <strong>of</strong> these<br />

trials are given below.<br />

Summary <strong>of</strong> field trials conducted with NPV <strong>of</strong> redheaded pine sawfly, Neodiprion<br />

lecontei (Fitch), between 1976 and 1980<br />

Ontario<br />

No. <strong>of</strong> No. <strong>of</strong><br />

Area planta- plantatreated<br />

tions tions<br />

Year (ha) from air from<br />

ground<br />

1976 43 3<br />

1977 52 3<br />

1978 26 2<br />

1979 34 4<br />

1980 540 8 88<br />

Ground spray trials in Quebec in 1970<br />

Area<br />

treated<br />

(ha)<br />

700<br />

330<br />

21<br />

Quebec<br />

No. <strong>of</strong><br />

plantations<br />

from air<br />

36<br />

42<br />

No. <strong>of</strong><br />

plantations<br />

from<br />

ground<br />

Purified NPV at a concentration <strong>of</strong> 1.3 x 10" PIBII and lyophilized NPV-infected larvae<br />

(with an undetermined PIB content) at 30, 60, 130, and 260 mgll were sprayed on thirdinstar<br />

larvae using a backpack mist blower. First mortality was observed 14 days after<br />

spraying and by 48 days 100% mortality was observed in the area treated with the<br />

lyophilized NPV-infected larval material, and 94.9 and 97.4% mortality in two areas<br />

treated with the purified virus (Anon. 1970).<br />

1<br />

1<br />

12


Aerial spray trials in Ontario in 1976<br />

Neodiprioll lecollId (Fitch). 325<br />

Three red pine plantations with a total area <strong>of</strong> 43.2 ha were sprayed with three different<br />

dosages <strong>of</strong> virus, 1.25 x Hr, 3.75 x 10", and 6.75 x to' PIBlha, at 9.4 Uha, using a fixedwing<br />

aircraft with boom and nozzle equipment, when most larvae were in the second<br />

instar. The formulation contained 60 gil !MC 9Q.OO1 (International Minerals Corporation)<br />

sunlight protectant. Pre-spray colony counts for the three plots were 59, 138, and 173<br />

colonies/50 trees, respectively, and 119 colonies/50 trees in an untreated check area.<br />

Seventeen days after spraying, the percentage <strong>of</strong> larvae infected with virus in the treated<br />

plots was 92, 96, and 98%, respectively, and 4 % in the check area. When the last colony<br />

count was made at 22 days, 7,46, and 1 colony/50 trees were recorded in the treated plots<br />

and 115/50 trees in the check plot. There was a good level <strong>of</strong> foliage protection in the<br />

treated areas (Kaupp & Cunningham 1977).<br />

Aerial spray trials in Ontario in 1977<br />

Two red pine and one jack pine plantation with a combined area <strong>of</strong> 48 ha were treated at a<br />

dosage <strong>of</strong> 5.5 x 10" PlBlha at 9.4 Uha. A fixed-wing aircraft with boom and nozzle<br />

equipment was used to spray second- and third-instar larvae on red pine, and a fixedwing<br />

aircraft with Micronair (Micronair4'i (Aerial) Ltd. U. K.) equipment to spray fourthinstar<br />

larvae on jack pine. The aqueous formulation contained 25% v/v molasses and 60<br />

gil IMC 90-001.<br />

In the jack pine plantation, 5 days after spraying the number <strong>of</strong> colonies/100 trees fell<br />

from 132 to 47 while the population in the untreated check plot remained constant. This<br />

rapid population decline prompted an investigation <strong>of</strong> the tank mix and it was found to<br />

contain 522JLglml <strong>of</strong> the chemical insecticide phosphamidon (Sundaram 1977 personal<br />

communication). By 26 days after spraying, the population density had fallen to 2<br />

colonies/100 trees.<br />

In the two red pine plantations, pre-spray counts <strong>of</strong> colonies/100 trees were 163 and<br />

54, and 217 in a check plot. Counts 26 days after spraying were 7 and 0 in the treated plots<br />

and 200 in the check area. The NPV also spread to an untreated plantation adjacent to<br />

these plantations and destroyed the redheaded pine sawfly population, but not before<br />

severe defoliation had occurred. Defoliation was light in all the treated areas, although it<br />

would have been heavy in the jack pine plantation without the unintentional contamination<br />

by phosphamidon in the aircraft spray tank (Kaupp et al. 1978).<br />

Aerial spray trials in Ontario in 1978<br />

Two red pine plantations with a total area <strong>of</strong>26 ha were treated when larvae were mainly<br />

in the second instar. The spray was applied with a fixed-wing aircraft equipped with a<br />

Micronair4'i spray delivery system. The dosage was 5 x 10" PIBlha in 9.4 Uha aqueous<br />

formulation containing 12.5% v/v molasses and 32 gil Shade® (Sandoz Inc.) (formerly<br />

IMC 90-001). Over 90% <strong>of</strong> sawfly colonies were diseased or dead 16 days after spraying<br />

and by 23 days no healthy colonies remained. Defoliation was lighter in treated than in<br />

check areas (de Groot et al. 1979).<br />

Aerial and ground spray trials in Quebec in 1978<br />

A large operation was undertaken and 37 red pine plantations covering 700 ha were<br />

treated. Two types <strong>of</strong> aerial spray equipment were used, a fixed-wing aircraft equipped<br />

with Micronair® and a helicopter equipped with Beecomist (Beeco Products Co., USA).<br />

In addition, one plantation was treated from the ground using backpack mist blowers.<br />

Dosage was 5 x 10' PIB/ha for aerial applications on second-instar larvae and 10 10 PIBlha


326 J. C.Cunningham and P. Dc Groot<br />

on third- and fourth-instar larvae for the ground application. Volume applied in all<br />

treatments was 9.4 l/ha. Some NPV was formulated in 25% v/v molasses and 30 gil<br />

Shade® and some was applied in water alone. Two application strategies were used, with<br />

some plantations getting total coverage and others partial coverage using widely spaced<br />

spray swaths. By 29 days after spraying, plantations sprayed by fixed-wing aircraft had<br />

an average mortality <strong>of</strong> 97.7% with total coverage and 72.5% with partial coverage.<br />

Plantations treated by helicopter had an average mortality <strong>of</strong> 99.4% with total coverage<br />

and 90.8% with partial coverage. The plantation treated with mist blowers had 98.4%<br />

mortality and untreated check plots had 29.5% mortality. The saving <strong>of</strong> foliage was<br />

satisfactory in all treatments except for the ground spray application where the application<br />

volume <strong>of</strong> 9.4 I/ha was too low and insect development was too far advanced (Desaulniers &<br />

Cunningham unpublished).<br />

Aerial spray trials in Ontario in 1979<br />

Red pine plantations with a total area <strong>of</strong> 33.6 ha were treated at a dosage <strong>of</strong> 5 x 10" PIBlha<br />

at 9.4 llha when larvae were in the first and second instars. A fixed-wing aircraft with<br />

Micronair® equipment was used. A comparic:on was made between a suspension <strong>of</strong><br />

lyophilized, NPV-infected larvae with Rhodamine B dye as a tracer and the same virus<br />

suspension formulated in 25% v/v molasses and 60 g.1 Shade®. There was no difference<br />

in the efficacy <strong>of</strong> the two treatments and it was concluded that it is unnecessary to add<br />

adjuvants to this virus. Both treatments gave 100% mortality, none <strong>of</strong> the larvae<br />

developed past the third instar and defoliation was minimal (Cunningham & de Groot<br />

unpublished).<br />

Aerial and ground spray trials in Quebec in 1979<br />

Forty-three red pine plantations with a total area <strong>of</strong> 330 ha were treated; 323 ha were<br />

sprayed from the air and 7 ha from the ground. Two helicopters were used for the aerial<br />

spray, one equipped with Beecomist411 units and one with boom and nozzle. From the air,<br />

application rate was 9.4 l/ha and dosage was 5 x 10" PIBlha. Backpack mist blowers<br />

were used for the ground spray application; volume sprayed was 18.8 l/ha and dosage<br />

was 10 10 PIBlha. Spraying was carried out when 80-90% <strong>of</strong> the eggs were hatched. The<br />

pre-spray population densities were 63 colonies/50 trees in the area treated with Beecomist®,<br />

110/50 trees in the area treated with boom and nozzle, 110/50 trees in the area treated<br />

with mist blowers and 22150 trees in the check area. At 35 days after spraying, mortality<br />

in these areas was 87, 98, 94 and 7%, respectively. Defoliation in the treated plantations<br />

was practically nil and only light defoliation was recorded in the check plots because <strong>of</strong><br />

the low sawfly population density (Bordeleau unpublished).<br />

Aerial and ground spray trials in Ontario in 1980<br />

Red pine plantations with a total area <strong>of</strong> 539.8 ha were treated. Experimental aerial spray<br />

trials were conducted on four plantations with a total area <strong>of</strong> 33 ha. A fixed-wing aircraft<br />

with Micronair® equipment was used and dosage on all plantations was 5 x 1()O PIBlha.<br />

Highly purified virus at 9.4 l/ha emitted volume was compared to lyophilized, NPVinfected<br />

larval material at the same volume. A comparison was also made with reduced<br />

emitted volumes <strong>of</strong> lyophilized larval material at 4.7 and 2.4 l/ha. Applications were<br />

made on first- and second-instar larvae. There was no difference in effectiveness<br />

between the crude and purified material or between the three emitted volumes. By 31<br />

days after spraying, 100% mortality was achieved in all four treatments.


Table 85<br />

Neodipriotl {ecomei (Fitch). 327<br />

Semi-operational aerial spray trials were conducted on a further four plantations with<br />

a total area <strong>of</strong>70 ha. Lyophilized NPV-infected larval material was used at 5 x 10' PIBlha<br />

and the emitted volume was 9.4l1ha. Larvae were mainly in the first and second instars,<br />

with a few unhatched egg clusters present, at the time <strong>of</strong> application. When the final<br />

survey was made 28 days after spraying, mortality ranged from 82 to 99%.<br />

Ground spray trials were conducted by Ontario Ministry <strong>of</strong> Natural Resources staff in<br />

five districts, and 88 plantations with a total area <strong>of</strong> 436.8 ha were treated entirely or<br />

partially over a 47-day period. Insect development ranged from first and second instar in<br />

the early treatments to fourth and fifth instar in the later ones (Cunningham & de Groot<br />

unpublished).<br />

Ground spray trials in Quebec in 1980<br />

A mist blower was used to treat 12 red pine plantations when larvae were in the first and<br />

second instars with 90% <strong>of</strong> the eggs hatched. The dosage was 5 x 10' PIBlha and two<br />

application volumes were tested, 18.7 and 93.6I1ha. Better results were obtained with<br />

the lower volume and 100% mortality was achieved 20 days after spraying. Thirty-five<br />

days after spraying, 97% mortality was reached with the higher volume. Defoliation in<br />

the treated plantations was negligible because <strong>of</strong> comparatively low larval population<br />

densities (an average <strong>of</strong> 21 colonieS/25 trees on the high volume application and 8<br />

colonies/l00 trees on the low volume application) combined with the effectiveness <strong>of</strong> the<br />

treatments (Bordeleau unpublished).<br />

Surveys <strong>of</strong> plantations aerially sprayed in Ontario in years following the year <strong>of</strong> application<br />

In order to assess the durability <strong>of</strong> the virus treatments, surveys have been conducted<br />

annually on all the plantations aerially sprayed experimentally in Ontario, which by 1980<br />

numbered 15 (de Groot et al. 1979, Cunningham & de Groot unpublished). A summary<br />

<strong>of</strong> these surveys is shown in Table 85. Although not all these plantations remained<br />

Survey <strong>of</strong> plantations treated with NPV <strong>of</strong> redheaded pine sawfly, Neodiprion leeontei<br />

(Fitch), in years following the year <strong>of</strong> application<br />

Number <strong>of</strong> colonieS/l00 trees<br />

Pre-spray<br />

Year Plot estimate 1977 1978 1979 1980 1981<br />

1976 1 118 0 0 0 0 0<br />

2 176 0 0 0 0 0.75<br />

3 346 0 0 0 0 0<br />

1977 1 163 0 0 0 0.75<br />

2 81 0 0 1.75 0<br />

3 132 0 0 0 0.5<br />

1978 255 0.5 0 0<br />

2 174 2.5 1.25 2.0<br />

1979 1 124 0.25 0<br />

2 77 0 0.25<br />

3 5 0 0<br />

1980 1 52 0<br />

2 173 0<br />

3 225 0<br />

4 25 0


Blank Page<br />

330


Pest Status<br />

Background<br />

Chapter 58<br />

Neodiprion sertifer (Ge<strong>of</strong>froy), European<br />

Pine Sawfly (Hymenoptera: Diprionidae)<br />

K.J. GRIFFITHS, J.e. CUNNINGHAM and I.S. OTVOS<br />

The European pine sawfly, Neodiprion serlifer (Ge<strong>of</strong>fr.), was first recorded in North<br />

America in New Jersey in 1925 (Schaffner 1939) and in <strong>Canada</strong> near Windsor, Ontario,<br />

in 1939 (Raizenne 1957). In Ontario it dispersed steadily until, by 1968, it was found<br />

throughout Ontario south and west <strong>of</strong> a line from Victoria Harbour to Belleville (Fig. 17).<br />

Transport <strong>of</strong> infested nursery stock resulted in isolated populations <strong>of</strong> N. serlifer well<br />

beyond this area - on Manitoulin Island, first recorded in 1966 (Sippell el al. 1966) and<br />

in Sault Ste. Marie and North Bay, first recorded in 1968 (Sippell el al. 1969).<br />

In the decade since 1968 the area occupied by N. serlifer has continued to expand,<br />

moving slightly to the north near Georgian Bay, but mainly to the east along the St.<br />

Lawrence River, until by 1978 the continuous distribution extended almost to the<br />

Quebec border along the St. Lawrence and as far north as Ottawa (Fig. 17) (Lindquist &<br />

Miller 1979).<br />

The isolated infestations on Manitoulin Island and at Sault Ste. Marie continued<br />

throughout the period 1968-78, but no N. senifer have been found in North Bay since<br />

1969. N. sertifer larvae were collected for the first time at several locations in Ottawa in<br />

1969 and by 1973 they were also found in the surrounding areas. They apparently<br />

remained there as Ottawa is now in an area <strong>of</strong> continuous distribution. An isolated light<br />

infestation was found north <strong>of</strong> Thessalon, Ontario, in 1974. The area was treated from<br />

the ground with nuclear polyhedrosis virus (NPV) in 1975 and 1976, and there have been<br />

no further recoveries <strong>of</strong> N. serliler.<br />

In 1970 there were several heavy infestations in southwestern Ontario, where the<br />

insect has been present the longest. Infestations causing moderate to severe defoliation<br />

continued in this area until 1972. A decrease in numbers was noted in 1973, and this<br />

decrease continued until by 1977 generally low populations were recorded throughout<br />

the area.<br />

In 1974 N. sertiler was recorded for the first time in <strong>Canada</strong> beyond the borders <strong>of</strong><br />

Ontario. A single colony was recovered from Scots pine, Pinus sylvestris L., in Charlesbourg,<br />

Quebec, a few kilometres northeast <strong>of</strong> Quebec City (Martineau & Lavallee 1975), and a<br />

few colonies were recovered from Scots pine at Windsor Lake near St. John's, Newfoundland<br />

(Clark & Singh 1975). No further recoveries were made in Quebec, but N. serlifer<br />

has persisted and spread in Newfoundland. It was reported on ornamental pines Pinus<br />

spp. and in the few pine plantations within a radius <strong>of</strong> 15 km <strong>of</strong> St. lohn's by Otvos &<br />

Griffiths (1979). It was established by rearing some larvae that no parasitoids were<br />

present. The first discovery <strong>of</strong> N. serlifer in Nova Scotia was made in 1980, when it was<br />

obtained from ornamental pines at Little Harbour, Pictou County, and Truro, Colchester<br />

County (Magasi 1981).<br />

Studies <strong>of</strong> the population dynamics <strong>of</strong> N. sertifer carried out at the Great Lakes Forest<br />

Research Centre, Sault Ste. Marie, Ontario, had started in 1960 and were continued until<br />

1972. An overview <strong>of</strong> this work is given in Lyons el al. (1971) and Lyons (1977a). In<br />

brief, outbreaks in southern Ontario typically occur in young plantations where few<br />

natural control agents are present. Density peaks in 4-6 years, then declines to a low<br />

331


Neot!i/Jrio" satifer (Ge<strong>of</strong>froy), 333<br />

level and does not generally increase to outbreak levels again. Eventual decline <strong>of</strong> an<br />

outbreak was related to starvation <strong>of</strong> late-feeding larvae, parasitism <strong>of</strong> pre-spinning<br />

larvae by Exenlerus spp. and predation and parasitism <strong>of</strong> cocooned larvae. When<br />

outbreaks occur in stands <strong>of</strong> larger trees, they last longer and are associated with varying<br />

amounts <strong>of</strong> prolonged diapause (Lyons 1977a).<br />

Until 1958 the recorded insect parasitoid complex <strong>of</strong> N. sertifer consisted <strong>of</strong> 14 native<br />

primary parasitoids, 3 native hyperparasitoids, and 4 established European parasitoids.<br />

In the following decade another 14 indigenous parasitoids were recovered; recovery was<br />

made <strong>of</strong> Drino bolremica Mesn., another <strong>of</strong> the European species introduced during the<br />

1930s (McGugan & CoppeI1962); and one new exotic species, Lophyroplectus lutealor<br />

Thunb., was established (Griffiths & Lyons 1968). In the period under review only two<br />

species have been added to the parasitoid list. One <strong>of</strong> these, Exenterus affinis Roh., a<br />

native ichneumonid, was recovered from N. sertifer as early as 1964, but was not<br />

separated from the other species <strong>of</strong> Exenlerus attacking this host until 1977. It is<br />

recovered infrequently (Lyons 1977b). The second species is Exenterus abruptorius<br />

(Thunb.), a European parasitoid <strong>of</strong> N. sertifer, introduced in the 1930s but not recovered<br />

since the initial release. Lyons (1977b) recovered two specimens in a cocoon collection<br />

made in Dufferin County in May, 1972. Since then five more E. abruptorius from four<br />

other localities have been identified after examination <strong>of</strong> over 6 000 pinned specimens <strong>of</strong><br />

Exenterus obtained from N. ser/ifer collections throughout southwestern Ontario since<br />

1952. Further discussion <strong>of</strong> this species is given in the following section. Another<br />

recovery <strong>of</strong> the rare exotic species D. bohemiea was recorded in 1972 from a collection<br />

made at Gore Bay on Manitoulin Island.<br />

Dispersal <strong>of</strong> L. IUleator from the releases made in 1962 and 1964 (Griffiths & Lyons<br />

1968) has been followed closely. By 1969,7 years afterthe 1962 release in Grey County,<br />

recoveries had been made up to 19.3 km east <strong>of</strong> the release point. However, by 1971, 7<br />

years after the 1964 Norfolk County release, the greatest dispersal was only 3.4 km<br />

(Griffiths 1973). Rose (1976) reported that by 1975 L.luteatorhad spread throughout the<br />

whole <strong>of</strong> the N. sertifer distribution area east and north <strong>of</strong> the Grey County release<br />

point. It is possible, however, that the northern dispersal into the Bruce Peninsula was<br />

the result <strong>of</strong> a release made near Hepworth, Ontario, in 1970 (Griffiths & Lyons 1980).<br />

Rose noted that L. luteator had spread 305 km in an easterly direction in the 13 years<br />

since its release. There had been little or no movement south or west from the 1962<br />

release point and little further dispersal from the 1964 release in Norfolk County.<br />

Investigations on the biology <strong>of</strong> the principal parasitoids <strong>of</strong> N. sertifer and the<br />

interactions between them have continued during the last decade. It was found that<br />

females <strong>of</strong> Pleolophus basizonus (Grav.), which oviposit in cocooned pre-pupae, show<br />

a strong rejection <strong>of</strong> hosts containing larvae or pupae <strong>of</strong> its own species, although they<br />

do not discriminate between unattacked hosts and those containing eggs <strong>of</strong> its own<br />

species (Griffiths 1972). Also, it was found that P. basizonus females show a strong<br />

avoidance <strong>of</strong> cocoons in which there are eggs, larvae, or pre-pupae <strong>of</strong> the larval<br />

parasitoid L. lutealor (Griffiths 1976) and avoid cocoons containing developing E.<br />

abruptorius and Dahlbominus fuscipennis (Zelt.). On the other hand, D. fuscipennis<br />

does not avoid hosts containing P. basizonus or E. abruptorius and is usually successful<br />

in competition with them. Exenterus abruptorius, E. nigrifrons (formerly E. canadensis<br />

Prov.) and E. amielorius (Panz.) all superparasitize and none avoids attacking hosts<br />

containing eggs <strong>of</strong> either <strong>of</strong> the other two species. Neither E. abruptorius nor E.<br />

nigrifrons has an advantage over the other when competing on the host, but E. amictorius<br />

is usually successful when competing with the other species. This information on<br />

Exenterus spp .• plus data on the three species obtained in insectary experiments over 6<br />

years, was incorporated into a simulation model to investigate the interactions <strong>of</strong> the<br />

three parasitoids and their host. More recently a more sophisticated simulation model <strong>of</strong><br />

N. sertifer and the living and non-living factors that influence it has been developed. It


Table 86<br />

Open releases and recoveries<br />

Neodiprion sertiler (Ge<strong>of</strong>fr.)<br />

Neodipriol/ satirer (Ge<strong>of</strong>froy). 335<br />

<strong>of</strong> parasitoids against the European pine sawfly,<br />

Number Year <strong>of</strong><br />

Species and province Year Origin released recovery<br />

Dipriocampe diprioni Ferriere<br />

Ontario 1972 Austria 78<br />

1974 Germany 183 1975<br />

1978 Italy<br />

{<br />

33<br />

Italy<br />

1979 Switzerland<br />

and Austria<br />

272<br />

Exenterus abruptorius Thnb.<br />

Ontario 1972 Austria 648<br />

1973 Austria 1021<br />

1979 Austria 967<br />

1979 Italy 1035<br />

1980 Finland 235<br />

1980 Austria 90<br />

Newfoundland 1980 Finland 75<br />

Lophyroplectus lWeator (Thnb.)<br />

Ontario 1970 Ontario 80 1970<br />

1972 Ontario 2263 1972<br />

Newfoundland 1978 Ontario 306<br />

1979 Ontario 38<br />

1980 Ontario 467<br />

1980 Austria 50<br />

1980 Finland 1<br />

Pleolophus basizonus (Grav.)<br />

Newfoundland 1977 Ontario 1007 1977<br />

1978 Ontario 2104<br />

1979 Ontario 1007<br />

1980 Finland 25<br />

infested with N. sertiler near Williamsford, Ontario. Eggs <strong>of</strong> N. sertiler were collected<br />

in the release area and examined in the spring <strong>of</strong> 1962 and spring egg collections were<br />

made annually until 1968 in a population study plot 0.8 km away. No recoveries <strong>of</strong> the<br />

parasitoid were obtained.<br />

Because the egg stage <strong>of</strong> N. seniler in <strong>Canada</strong> was essentially an untapped resource<br />

for biological control. it was decided that more study on European egg parasitoids was<br />

justified in spite <strong>of</strong> earlier failures. The work was carried out by CIBC European<br />

Laboratory scientists. As a result <strong>of</strong> this work (Pschorn-Walcher & Eichhorn 1971,<br />

1973) it was found that D. diprioni had a univoltine strain adapted to N. seniler and a<br />

morphologically similar bivoltine strain which occurs rarely in N. sertiler. Earlier<br />

failures may have been the result <strong>of</strong> using the wrong strain and further work on<br />

introduction. using the univoltine strain, was justified. This work started in 1971 with a<br />

preliminary test which indicated that D. diprioni had attacked Canadian N. sertiler and<br />

had overwintered successfully outdoors near Chatsworth, Ontario.<br />

A release <strong>of</strong> 32 male and 46 female adults was made in Dufferin County Forest, 5<br />

September 1972 (Table 86). but this area was sprayed inadvertently with a chemical


336 K. J. Griffiths. J. C. Cunningham and I. S. Dtvos<br />

Table 87<br />

Exenterus abruptDrius<br />

(Thunb.)<br />

(Hymenoptera:<br />

Ichneumonidae)<br />

insecticide in the summer <strong>of</strong> 1973 and the N. sertiter population there was destroyed.<br />

Another release was made, in Grey Country, 6 September 1974, <strong>of</strong> 29 males and 154<br />

females (Table 86). There was 1.3% parasitism in 87 egg clusters collected at the release<br />

site in the spring <strong>of</strong> 1975. Parasitoids were obtained, but in decreasing numbers in both<br />

1976 and 1977; none was obtained in the following 3 years (Table 87).<br />

Releases <strong>of</strong> D. diprioni have been made at two other localities, one in Simcoe County,<br />

consisting <strong>of</strong> 11 male and 22 female adults, on 16 September 1978, and the other, <strong>of</strong> 90<br />

male and 182 female adults, in Peterborough County, 28 August and 30 September 1979<br />

(Table 86). There have been no recoveries from either release (Table 87).<br />

Laboratory rearings <strong>of</strong> D. diprioni were started in 1974 using adults obtained from<br />

Switzerland and Germany. Work continued until 1977 using the <strong>of</strong>fspring <strong>of</strong> the original<br />

stock augmented with adults from field collections <strong>of</strong> N. sertifer eggs and further<br />

shipments from Europe (Table 88) but it was not possible to maintain stock beyond the<br />

third generation.<br />

Recovery <strong>of</strong> Dipriocampe diprioni Ferriere<br />

Number <strong>of</strong> Number <strong>of</strong><br />

clusters Number <strong>of</strong> egg clusters Number <strong>of</strong> Sex ratio<br />

Year collected eggs examined with attack eggs attacked (% females)<br />

Grey County, Ontario<br />

1975 87 6052 7 80 (1.3%) 53<br />

1976 57 3094 2 34 (1.1%) 50<br />

1977 86 3819 1 7 (0.2%) 71<br />

1978 6 430 0 0<br />

1979 144 ? 0 0<br />

1980 28 ? 0 0<br />

Simcoe County, Ontario<br />

1979 52 ? 0 0<br />

1980 48 ? 0 0<br />

Peterborough County, Ontario<br />

1980 47 ? 0 0<br />

Since the recovery <strong>of</strong> a small number <strong>of</strong> adults <strong>of</strong> this species in southwestern Ontario in<br />

the spring <strong>of</strong> 1972, noted earlier, there has been renewed interest in trying to establish it<br />

at higher densities and in determining why it has remained so rare since its introduction<br />

in the period from 1936 to 1949 (McGugan & Coppel 1962). A number <strong>of</strong> well-timed<br />

releases have been made in Ontario (Table 86) but because N. sertiter popUlations are<br />

generally low, releases could only be made in areas <strong>of</strong> low host density. At these<br />

population levels large cocoon collections would have been necessary to determine if<br />

establishment had occurred, and the cost made sampling unpractical.<br />

A release <strong>of</strong> 43 male and 32 female adults obtained from Finland was made in SI.<br />

John's, Newfoundland, in July 1980 (Table 86).<br />

Studies on the biology <strong>of</strong> this species were started with insectary experiments in 1969<br />

and 1970 and continued with laboratory experiments from 1978 to 1980, using E.<br />

abruptorius adults from various European locations (Table 88). Although analysis <strong>of</strong><br />

results is not complete it is already clear that no single physical factor has been limiting<br />

the numbers <strong>of</strong> this species in Ontario.


PfeoIopbus ba9zonus<br />

(Grav.)<br />

(Hymenoptera:<br />

Ichneumonidae)<br />

Pathogens<br />

Table 88<br />

Neotiipriol/ .f(°r/iler (Ge<strong>of</strong>froy), 337<br />

Laboratory and field cane studies <strong>of</strong> parasitoids against the European pine sawfly<br />

Neodiprion seniler (Geo r.)<br />

Species and province Year Origin Number<br />

Dipriocampe diprioni Ferriere<br />

Ontario 1970 Austria 196<br />

1971 Austria 228<br />

1974 Germany 73<br />

1975 Austria 19<br />

1976 Switzerland 136<br />

1977 Switzerland 72<br />

£Xenlerus abruplorius Thnb. 1969 Austria 579<br />

Ontario 1970 Austria 244<br />

1978 Austria 242<br />

1978<br />

1978<br />

Germany<br />

Germany { Austria<br />

64<br />

75<br />

1978 Italy 211<br />

1979 Finland 139<br />

As stated earlier, shipments <strong>of</strong> parasitoids to Newfoundland were started soon after it<br />

was known that N. senifer was established there. The first species to be shipped was P.<br />

basizonus, a multivoltine parasitoid <strong>of</strong> cocooned pre-pupae successfully introduced from<br />

Europe in the 19305 (McGugan & Coppell962). N. seniler cocoons containing developing P.<br />

basizonus were shipped to Newfoundland in the summer <strong>of</strong> 1977. They were obtained<br />

from a rearing programme conducted in Sault Ste. Marie, Ontario, during the winter<br />

preceding shipment, and were reared in St. John's. Emerging parasitoids were released<br />

twice weekly between 21 July and 15 August at Windsor Lake, 9.7 km from St. John's. A<br />

total <strong>of</strong> 631 males and 376 females was released in an infested plantation <strong>of</strong> mixed jack<br />

pine and Scots pine (Otvos & Griffiths 1979) (Table 86). Approximately 130 adult P.<br />

basizonus were recovered from 600 N. seniler cocoons "planted" in the release area in<br />

the summer <strong>of</strong> 1977, indicating that the released adults had attacked successfully in<br />

the field (Otvos & Griffiths 1979). Further releases <strong>of</strong> large numbers <strong>of</strong> this parasitoid<br />

obtained from rearings in Sault Ste. Marie were made in 1978 and 1979. A small number<br />

<strong>of</strong> adults from Finland was also released in 1980 (Table 86). There have been no further<br />

recoveries <strong>of</strong> adults <strong>of</strong> this species from field-collected cocoons to date. Monitoring <strong>of</strong><br />

this release is continuing.<br />

In Ontario, European pine sawfly NPV was extensively used in the 19505 and 1960s by<br />

Christmas tree growers and provincial government forestry <strong>of</strong>ficials.<br />

There are several methods <strong>of</strong> producing this virus in host insect larvae, the simplest<br />

and cheapest being to do so in the field. Plantations with a suitable insect population<br />

density are found and fourth-instar larvae are sprayed using a mist blower to disseminate<br />

NPV at a concentration <strong>of</strong> 10" PIB/ml. The first dead larvae are found about 7 days after<br />

spraying and colonies <strong>of</strong> dead and dying larvae are collected daily until about 14 days<br />

after spraying. Larvae are removed from the foliage, frozen, lyophilized, ground to a<br />

fine powder, and the concentration <strong>of</strong> PIB per gramme determined. About lOS PIB are<br />

obtained from one dead larva, so, using a dosage <strong>of</strong> 5 x 10" PIBlha. about 50 virus-killed<br />

larvae are required. A virus treatment is therefore very economical. Between 1970 and<br />

1975 a considerable quantity <strong>of</strong> this virus was produced by Ontario Ministry <strong>of</strong> Natural


Recommendations<br />

Neocliprioll sertifer (Ge<strong>of</strong>froy). 339<br />

There is no doubt that European pine sawfly NPV is a highly efficacious biological<br />

control agent when applied on early instar larvae, and treatments continue to give control<br />

in the years following application. The longevity <strong>of</strong> this control depends on such factors as<br />

the initial population density <strong>of</strong> the host and the size <strong>of</strong> the trees, but infectious PIBs<br />

remain in the environment for at least 3 years following a virus epizootic (Kaupp 1981).<br />

Given this respite from defoliation, trees grow considerably and are much less susceptible<br />

to damage if further European pine sawfly outbreaks occur.<br />

To date, there are no viruses registered for operational use in <strong>Canada</strong>. In the USA, a<br />

petition has been prepared for registration <strong>of</strong> European pine sawfly virus under the name<br />

"Neochek-S". Extensive animal safety testing has been conducted by the Americans, but<br />

it is not known if these data will be acceptable to Canadian authorities. With the limited<br />

areas <strong>of</strong> European pine sawfly infestation currently reported, registration <strong>of</strong> its NPV in<br />

<strong>Canada</strong> is not <strong>of</strong> high priority in comparison to viruses <strong>of</strong> certain other insect pests. Even<br />

if European pine sawfly NPV is eventually registered in <strong>Canada</strong>, there is little commercial<br />

incentive for production and marketing owing to its limited and specialized use. Virus<br />

production and distribution, therefore, will almost certainly remain in the hands <strong>of</strong><br />

federal and provincial governments.<br />

In view <strong>of</strong> the successful establishment <strong>of</strong> L. lutealor in five areas <strong>of</strong> Ontario, all following<br />

a single release, the failure to establish this species in Newfoundland following three<br />

releases in 3 successive years suggests a fundamental difference between the situations in<br />

Ontario and Newfoundland. We recommend that no more releases be made there after<br />

the 1981 release until further study reveals success or the reasons for failure.<br />

It was pointed out by Griffithsetal. (1971) that there are 62 species <strong>of</strong> native parasitoids<br />

known to attack other pine-feeding diprionids in Ontario that had not been recovered<br />

from N. serlifer. We have not recovered any <strong>of</strong> them in the limited sampling in the last<br />

decade in southern Ontario. It is possible that some <strong>of</strong> them are now attacking N. serlifer<br />

in northern Ontario and we recommend that sampling be done to determine if this is so.<br />

No more releases <strong>of</strong> D. diprioni or E. abruptorills should be made. However, sampling<br />

to determine whether the former species has become established can be continued with<br />

little cost for several years. Collections to determine the presence <strong>of</strong> E. abruplorius<br />

should only be undertaken when they can provide adequate numbers <strong>of</strong> hosts in order to<br />

justify the work that the processing <strong>of</strong> collections requires.<br />

As stated earlier, the operational use <strong>of</strong>NPV at Sandbanks Provincial Park in Ontario<br />

was considered to be an unqualified success. However, after assessing aerial applications<br />

<strong>of</strong> European pine sawfly NPV and evaluating reports <strong>of</strong> operations conducted outside<br />

<strong>Canada</strong>, two recommendations can be made. Firstly. the dosage <strong>of</strong> virus applied is about<br />

10 times more than required for satisfactory control. A dosage <strong>of</strong> 5 x Hf PIBlha is<br />

adequate. Secondly, adjuvants in the tank mix are unnecessary and virus in water alone<br />

will suffice.<br />

Although it is not within the discipline <strong>of</strong> biological control, we are concerned about the<br />

dispersal <strong>of</strong> the European pine sawfly beyond its present boundaries in <strong>Canada</strong>. Sullivan<br />

(1965) pointed out that the overwintering eggs <strong>of</strong> this species can survive a temperature <strong>of</strong><br />

-26°C and that, with conditioning and selection, they may be able to withstand<br />

temperatures even lower. Thus the possibility <strong>of</strong> this pest spreading throughout the range<br />

<strong>of</strong> its widely distributed hosts, Scots pine, jack pine, and red pine, is great unless efforts<br />

are made to prevent it. We recommend that some attempt be made to fumigate stock<br />

before transport to reduce this possibility, using techniques outlined by Monro & Kirby<br />

(1963).


340 K. J. Griffiths, J. C. Cunningham and I. S. Otvos<br />

Literature Cited<br />

Bird. F.T. (1950) The dissemination and propagation <strong>of</strong> a virus disease affecting European pine sawny. Neodiprion sertifer (Ge<strong>of</strong>f.).<br />

Canadian Department <strong>of</strong> Agriculture Bi·monthly Progress Report 6.2-3.<br />

Clark. R.C.; Singh. P. (1975) Newfoundland region. Canadian Forestry Service, Annual Report <strong>of</strong> the Forest Insect and Disease Survey<br />

1974.<br />

Cunningham. J.C.; Kaupp, W.J.; McPhee, J.R.; Sippell. W.L.; Barnes. C.A. (1975) Aerial application <strong>of</strong> a nuclear polyhedrosis virus to<br />

control European pine sawny. Canadian Forestry Service Bi·monthly Research Notes 31.39-40.<br />

Cunningham. J.C.; Entwistle. P.F. (1981) Control <strong>of</strong> sllwnies by baculovirus. In: Burges. D. (Ed.) Microbial control <strong>of</strong> pests and plant<br />

diseases. 1970-1980. London; Academic Press.<br />

Grirfiths, K.J. (1972) Diserimination between parasitized and unparasitized hosts by Pleolophus basizonu.r (Hymenoptera: Ichneumonidae).<br />

Proceedings <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> Ontario \02.83-91.<br />

Grirfiths, K.J. (1973) The dispersal <strong>of</strong> the introduced parasitoid Lophyroplectus luteator (Hymenoptera: Ichneumonidae) following its<br />

release against Neodiprion sertifer (Hymenoptera: Diprionidae) in Ontario. Canadian Entomologist<br />

\05.833-836.<br />

Griffiths, K.J. (1976) Interactions between Pleolophus basizonus and Lophyropleaus IUleator (Hymenoptera: Ichneumonidae). two parasitoids <strong>of</strong><br />

Neodiprion sertifer (Hymenoptera: Diprionidae). Entomophaga 21.13-17.<br />

Grirfiths, K.J.; Lyons. L.A. (1968) The establishment <strong>of</strong> Lophyroplectus IuJeaJor (Hymenoptera: Ichneumonidae) on Neodiprion serrife,<br />

(Hymenoptera: Diprionidae) in Ontario. CaruuJUm Entomologist 100.1095-1009.<br />

Griffiths. K.l.; Lyons. L.A. (1980) Funher rcleru;es <strong>of</strong> LophyropleclUJi luteator (Hymenoptera: Ichneumonidac). an introduced parasite <strong>of</strong><br />

Neodiprion smifer (Hymenoptera: Diprionidae). Canadian Entomologist 112,421-426<br />

Griffiths. K.1.; Rose. A.H.; Bird. F.T. (1971) Neodiprion serrifer (Ge<strong>of</strong>f.). European pine sawfly (Hymenoptera: Diprionidae). In: Biological<br />

control programmes against insects and weeds in <strong>Canada</strong> 1959- 1968. Commonwtalth InstiJute <strong>of</strong> Biologiazl<br />

Control Technical Communication 4.150-162.<br />

Kaupp. W.J. (1981) Studies on the ecology <strong>of</strong> the nuclear polyhedrosis virus <strong>of</strong> the European pine sawny. Neodiprion sertife, (Ge<strong>of</strong>f.). Ph.D.<br />

Thesis. Oxford University. U.K.<br />

Kraemer. M.E.; Coppel, H.C.; Hall, D.J. (1979) The establishment <strong>of</strong> Lophyroplectus luteator, a larval parasitoid <strong>of</strong> the European pine<br />

sawfly. Neodiprion serrifer. in Wisconsin. University <strong>of</strong> Wisconsin FOlTSt RestJ1rch Note 225,3 pp.<br />

Lindquist, O.H.; Miller, W.1. (1979) The pine sawny that carne to dinner. Your Forests 12(1).\0- I 1.<br />

Lyons, L.A. (1977a) On the population dynamics <strong>of</strong> Neodiprioll sawflies. In: Kulman, H.M.; Chiang. H.C. (Eds.). Insect ecology - papers<br />

presented in the A.C. Hodson Lectures. University <strong>of</strong> Millnesota Agriculture Experiment Statioll Technical<br />

Bulletin 310,48-55.<br />

Lyons. L.A. (1977b) Parasitism <strong>of</strong> Neodiprioll sertifer (Hymenoptera: Diprionidae) by Exenlerus spp. (Hymenoptera: Ichneumonidae) in<br />

Ontario, 1962-1972, with notes on the parasites. Canadian Entomologist 109,555-564.<br />

Lyons. L.A.; Sullivan, C.R.; Wallace. D.R.; Griffiths, K.J. (1971) Problems in the management <strong>of</strong> forest pest populations. Proceedings <strong>of</strong><br />

the Tall Timbers Conference on Ecological Animal Control 1971, pp. 129-140.<br />

Magasi. L.P. (1981) Forest pest conditions in thc Maritimes in 1980. Canadian Forestry Service, Fredericton, New Brunswick, Report<br />

M-X-118,35 pp.<br />

Manineau, R.; Lavallee. A. (1975) Quebec region. Canadian Forestry Service, Annual Report <strong>of</strong> the Forestlnseet and Disease Survey 1974,<br />

pp.37-56.<br />

McOugan, D.M.; Coppel, H.C. (1962) Biological control <strong>of</strong> forest insects. 1910-1958. In: A review <strong>of</strong> the biological control attempts against<br />

insects and weeds in <strong>Canada</strong>. Commonwealth Institute <strong>of</strong> Biological Control Technical Communication<br />

2.35-127.<br />

Monro, H.A.A.; Kirby, C.S. (1963) Fumigation <strong>of</strong> nursery slock as a possible means <strong>of</strong> retarding the spread <strong>of</strong> Ihe European pine sawfly.<br />

Canadian Department <strong>of</strong> Forestry, Entomology and Pathology Branch Bi-monthly Progress Report<br />

19(3).2.<br />

Otvos, I.S.; Griffiths. K.J. (1979) Cocoon parasitoid <strong>of</strong> the European pine sawfly introduced into Newfoundland from Ontario. Canadiall<br />

Forestry Service, Bi·monthly Research Notes 35(4),20-21.<br />

Pschorn·Walcher. H.; Eichhorn. O. (1971) Untersuchungen uber die Eiparasitien (Hym., Chalcidoidea) der rotgclben Kiefern-Buschhornblattwespe<br />

Neodiprion sertifer Georf. (Hym. Diprionidae). Anzeiger fur Schaedlingskunde, P[1anzellSchutz,<br />

Umweltschutz 44(7).97- 103.<br />

Pschorn·Walcher, H.; Eichhorn, O. (1973) Studies on the biology and ecology or the egg parasites (Hym.: Chalcidoidea) <strong>of</strong> the pine sawfly<br />

Neodiprion sertifer (Ge<strong>of</strong>f.) (Hym.: Diprionidae) in central Europe. Zeitschrift fur angewandte Entomologie<br />

74,286-318.<br />

Raizenne. H. (1957) Forest sawnies <strong>of</strong> southern Ontario and their parasitoids. Canadian Department <strong>of</strong> Agriculture, Division <strong>of</strong> Forest<br />

Biology Publication 1009,45 pp.<br />

Rose. A.H. (1976) The dispersal <strong>of</strong> Lophyroplectus luteator (Hymenoptera: Ichneumonidae) on Neodiprion sertifer (Hymenoptera:<br />

Diprionidae) in southern Ontario. Canadian Entomologist 108.\395-1398.<br />

Scharfner. J. V. (1939) Neodiprion smifer (Ge<strong>of</strong>r.) a pine sawfly accidentally introduced into New Jersey from Europe. lournal <strong>of</strong> Economic<br />

Entomology 32.887-888.<br />

Sippell, W.L.; Dance. B.W.; Rose. A.H. (1966) Canadian Forestry Service, Report <strong>of</strong> the Insect and Disease Survey 1965. pp. 44-65.<br />

Sippell, W.L.; Gross. H.L.; Rose. A.H. (1969) Canadian Forestry Service, Report <strong>of</strong> the Insect and Disease Survey 1968, pp. 54-66.<br />

Sullivan. C.F. (1965) Laboratory and field investigations on the ability <strong>of</strong> eggs <strong>of</strong> the European pine sawny. Neodiprion sertifer (Ge<strong>of</strong>froy) to<br />

withstand low winter temperatures. Canadian Entomologist 97.978-993.


Pest Status<br />

Background<br />

Releases and Recoveries<br />

Chapter 59 341<br />

Neodiprion swainei (Middleton), Swaine<br />

Jack Pine Sawfly (Hymenoptera:<br />

Diprionidae)<br />

R.J. FINNEGAN and W.A. SMIRNOFF<br />

At the time <strong>of</strong> the last review by McLeod & Smirn<strong>of</strong>f (1971), the population <strong>of</strong><br />

Neodiprion swainei (Middleton), Swaine jack pine sawfly, was approaching endemic<br />

levels over most <strong>of</strong> its range in eastern Ontario and southwestern Quebec, where it has<br />

occurred extensively since the early 1950s. This was thought to be due partly to the<br />

natural collapse <strong>of</strong> a prolonged infestation and partly to the application <strong>of</strong> chemical<br />

insecticides in 1965 (Mcleod 1968). Mcleod & Smirn<strong>of</strong>f (1971) had expected a<br />

recurrence <strong>of</strong> N. swainei in the 19705, but this did not materialize, probably due to earlier<br />

applications <strong>of</strong> chemical insecticides against both Swaine jack pine sawfly, N. swainei,<br />

(McNeil et al. 1979) and spruce budworm, Choristoneurafumiferana (Clem.). (Mcleod<br />

1975) in the same general area. The population remained low throughout the 1970s. with<br />

the occurrence <strong>of</strong> only a few localized light infestations in northern Ontario and western<br />

Quebec.<br />

Parasitoids, predators, and pathogens have all been introduced in attempts to control<br />

sawfly populations. McLeod & Smirn<strong>of</strong>f (1971) reported that "Before 1958, only small<br />

numbers <strong>of</strong> Ple%phus basizonlls (Grav.) and Drino bohemica Mesn. were released<br />

against Neodiprion swainei in a few localities in Ontario. However, massive releases <strong>of</strong><br />

these two parasitoids, as well as Exenterlls amictorius Panzer and Dahlbominus fuseipennis<br />

(Zett.) were also made in the early 1940s against Diprion hercyniae Htg. in both<br />

Ontario and Quebec. All four species have been recovered from N. swainei in a number<br />

<strong>of</strong> localities in Quebec since 1958 ... but none in Ontario. E. amictorius and P. basizonus<br />

were most abundant, D. fllseipennis was less common, and only one specimen <strong>of</strong> D.<br />

bohemica was found". A brief description <strong>of</strong> the interaction <strong>of</strong> each parasitoid with N.<br />

swainei was given. Mcleod & Smirn<strong>of</strong>f also reported on the discovery and successive<br />

recoveries <strong>of</strong> the Borrelina viral pathogen. The results <strong>of</strong> one experimental aerial<br />

dispersion <strong>of</strong> the virus in 1960 and a second in 1964 were reported as "excellent" and<br />

"not as good" respectively. The ecology <strong>of</strong> the virus was briefly considered, stating that<br />

certain predators, such as wasps and pentatomids act as vectors <strong>of</strong> the virus. In<br />

conclusion, McLeod & Smirn<strong>of</strong>f stated that the releases <strong>of</strong> E. amictorillS and P.<br />

basizonus seemed to have been beneficial, and that the use <strong>of</strong> virus as a control was<br />

promising.<br />

During the past decade, no further release <strong>of</strong> parasitoids, and few recoveries, were<br />

made. Nonetheless, considerable research was completed during this period. In anticipation<br />

<strong>of</strong> an impending resurgence <strong>of</strong> the N. swainei population in Quebec, two foreign<br />

species <strong>of</strong> predacious red wood ants, Formica lugubris Zett. and F. obscuripes Forel,


342 R. J. Finnegan and W. A. Smirn<strong>of</strong>f<br />

Parasitolds<br />

Predators<br />

were introduced into susceptible stands <strong>of</strong> jack pine Pinus banksiana Lamb., in the St.<br />

Maurice river valley. Observations were reported on other insect predators as well as on<br />

birds and small mammals. Studies on the virulence. formulation <strong>of</strong> sprays. and application <strong>of</strong><br />

pathogens were also continued.<br />

Mcleod (1971) examined the variation <strong>of</strong> N. swainei cocoon survival and found that<br />

only about 10% <strong>of</strong> the mortality was caused by parasitoids. Later. Mcleod (1972)<br />

compared host discrimination with density responses during oviposition by parasitoids<br />

<strong>of</strong> N. swainei; he observed that the ability <strong>of</strong> females to discriminate against hosts<br />

containing previously deposited parasitoid progeny was variable. Although there was<br />

no discrimination at the beginning <strong>of</strong> the spinning period <strong>of</strong> N. swainei. it was quickly<br />

acquired and persisted to the end <strong>of</strong> the spinning period.<br />

Price (1970a, 1970b, 1970c) showed that "Female parasitic insects in the genera<br />

Pleolophus, Ensasys, and Matrus (Hymenoptera: Ichneumonidae) search the ground<br />

cover for hosts and avoid areas they had already inspected" and that "females respond to<br />

their own trail odour, and recognition occurs also between conspecific, congeneric and<br />

intergeneric individuals". Later Price (1971) discussed the niche breadth and dominance<br />

<strong>of</strong> parasitic insects sharing the same host species. In 1972, he indicated that as a result <strong>of</strong><br />

recognition <strong>of</strong> repellent trail odours. females <strong>of</strong> P. basizonus are able to discriminate<br />

against parasitized hosts <strong>of</strong> N. swainei between narrow limits "set by the probability <strong>of</strong> a<br />

female finding an unparasitized host". and that they showed "mutual interference in egg<br />

laying at high parasitoid density" (Price 1972a). He also reported on the immediate and<br />

long-term effects <strong>of</strong> chemical insecticide application on insect parasites in jack pine<br />

stands in Quebec (Price 1972b); on the activity patterns and impact <strong>of</strong> paras ito ids on N.<br />

swainei (Price 1972c); and on the utilization <strong>of</strong> the same host by different insect parasites<br />

(Price 1972d).<br />

Finnegan (1971) reported on preliminary studies on the use <strong>of</strong> predacious ants to control<br />

forest pests in Quebec, and concluded that none <strong>of</strong> the species found was particularly<br />

promising. He further discussed (Finnegan 1974) both the harmful and beneficial aspects <strong>of</strong><br />

red wood ants as insect predators, the reason for their effectiveness, and recommendations<br />

for selecting desirable species for propagation. Finnegan (1975, 1977) later introduced<br />

two species <strong>of</strong> red wood ants, F. lugubris from central Europe and F. obscuripes from<br />

southern Manitoba, into a jack pine stand at Lac it la Chienne, about 70 km southwest <strong>of</strong><br />

La Tuque, Quebec, in the St. Maurice river watershed. The two species <strong>of</strong> ants belong to<br />

the F. rufa group and are aggressive predators <strong>of</strong> sawfly larvae, cocoons, and adults, as<br />

well as a variety <strong>of</strong> other defoliating insects (Cotti 1963, Finnegan 1978, McNeil et al.<br />

1978). These introductions were made in anticipation <strong>of</strong> a recurrence <strong>of</strong> N. swainei<br />

infestations in the late 1970s. When the initial introductions had become established,<br />

additional releases (from Europe and Manitoba) were made in subsequent years. Table<br />

89 shows the dates and approximate numbers <strong>of</strong> each ant species (workers and sexuals)<br />

released in the St. Maurice river watershed. The status <strong>of</strong> each species in 1981 is also given<br />

in Table 89, based on the number, average size, and insect vigour <strong>of</strong> the nests. Two<br />

secondary introductions <strong>of</strong> F. obscuripes (emanating from the first) were made in 1980 in<br />

the Gilardo Dam area, about 75 km west <strong>of</strong> La Tuque, Quebec.<br />

I1nytzky (1974) reported on the species <strong>of</strong> ants found in a young jack pine stand in<br />

Quebec, and on their distribution, population level, and predation <strong>of</strong> N. swainei. He<br />

studied 10 species <strong>of</strong> ants, <strong>of</strong> which F. fusca L. was the most abundant, and found that<br />

together with several other Formica species and Camponolus herculeanus (L.) they<br />

carried emerging and ovipositing adults, as well as cocoons, into their nests. He concluded that<br />

ants may be an important factor regulating N. swainei populations.


Pathogens<br />

Table 89<br />

Neodiprioll slI'ainei (Middleton), 343<br />

Open releases and recoveries <strong>of</strong> predacious red wood ants against Swaine jack pine sawfly,<br />

Neodiprion swainei (Middleton), in the St. Maurice river watershed in Quebec<br />

Approximate<br />

Status in 1981<br />

Ant<br />

species<br />

Date <strong>of</strong><br />

release<br />

number<br />

released Recovery<br />

Estimated<br />

increase<br />

Formica 1972 4 ()()() ()()() About 20 nests 15-fold<br />

obscuripes Forel established (with<br />

sexuals present)<br />

1975 1 ()()() ()()() About 75 nests 20-fold<br />

established (with<br />

sexuals present)<br />

F. lugubris Zett. 1973 1 ()()() ()()() Ten nests established<br />

(with few sexuals)<br />

1975 2 ()()() (added to previous ?<br />

queens nests)<br />

1976 5 ()()() (added to previous ?<br />

queens nests)<br />

Mcleod (1971) reported, in a preliminary analysis <strong>of</strong> the variation in cocoon survival<br />

<strong>of</strong> N. swainei, that small animal predators accounted for about 47% <strong>of</strong> the mortality <strong>of</strong><br />

cocoons, and insect predators for about 16%. Mcleod (1974) also found that populations <strong>of</strong><br />

resident breeding birds in jack pine stands in Quebec varied little from 1964 to 1973, but<br />

that the mole population showed a steady, statistically significant increase over the 9<br />

years <strong>of</strong> sawfly infestation.<br />

Tostowaryk (1971a) described in detail the life history and behaviour <strong>of</strong> Podisus<br />

modestus (Dallas), a predacious pentatomid, which preys on N. swainei larvae as well as<br />

on other defoliators. He gave details <strong>of</strong> the habits and abundance <strong>of</strong> the predator, and<br />

also discussed the relationship between parasitism and predation <strong>of</strong> diprionid sawflies<br />

(Tostowaryk 1971b). Experimental results <strong>of</strong> the functional response <strong>of</strong> P. modestus to<br />

densities <strong>of</strong> N. swainei and N. pratti banksianae Rohwer (Tostowaryk 1971c) showed<br />

that the response was two-fold: a domed curve when the sawfly larvae were active and<br />

relatively large with respect to the size <strong>of</strong> the predator, and a negatively accelerated<br />

curve when the sawfly larvae were freshly killed and relatively small with respect to the<br />

predator's size. He also presented a list <strong>of</strong> coleopterous predators <strong>of</strong> N. swainei<br />

(Tostowaryk 1972), and stated that it seemed these predators responded in a densityindependent<br />

manner to the prey. Tostowaryk (1973) continued his observations on<br />

carabids found in jack pine stands. He noted that the three most common species <strong>of</strong><br />

carabids preyed to a limited extent on cocoons <strong>of</strong> N. swainei.<br />

A highly pathogenic strain <strong>of</strong> a Borrelina species <strong>of</strong> virus was studied (Smim<strong>of</strong>f 1961a,<br />

Mcleod & Smim<strong>of</strong>f 1971) and a formulation suitable for aerial dispersion was developed<br />

(Smim<strong>of</strong>f 1964). Aerial dispersion <strong>of</strong> the virus provided excellent control <strong>of</strong> the insect<br />

(Smirn<strong>of</strong>f et al. 1962). The dosages used were 4.5 IIha and 3711ha, at a concentration <strong>of</strong><br />

2 x 1()6 polybedralml. Cold weather during and after treatment caused a reduction in<br />

larval mortality, but because <strong>of</strong> trans-ovum and trans-ovarial transmission <strong>of</strong> the<br />

disease from parent to progeny, there was a long-term control effect (Smim<strong>of</strong>f 1962).


Evaluation <strong>of</strong> Control Attempts<br />

Table 90<br />

Neodiprioll swainei (Middleton). 345<br />

and a decrease in free glycerol, potassium, and especially total lipids (energy resources),<br />

which were decreased by 50% (Smirn<strong>of</strong>f 1971b).<br />

Results showed that biochemical analysis can be used to determine the pathological<br />

state <strong>of</strong> an N. swainei population; to follow, metabolically, the evolution <strong>of</strong> a virosis in<br />

the insect, thus permitting a prognosis <strong>of</strong> its population; and to evaluate whether the<br />

population is expanding or decreasing. The metabolic study done during 1979 and 1980<br />

indicated that the N. swainei population <strong>of</strong> the Saguenay - Lac St-Jean region, Quebec,<br />

is expanding.<br />

The influence <strong>of</strong> urea fertilization on the physiology and growth <strong>of</strong> jack pine and on the<br />

behaviour <strong>of</strong> N. swainei (either healthy or virus-infected) was determined. It was found<br />

that a treatment with urea in natural forest stands increased the length and weight <strong>of</strong> jack<br />

pine needles proportionally to the amount <strong>of</strong> nitrogen assimilated. Three years after<br />

nitrogen fertilization, the ingestion <strong>of</strong> foliage by N. swainei was reduced by approximately<br />

50%, and the natural mortality <strong>of</strong> the tenthredinid was increased by 50%. It was found<br />

that the sensitivity <strong>of</strong> larvae to virus infection, as well as the mortality, increased. Also,<br />

modifications in the physiological state <strong>of</strong> jack pine needles provoked changes in insect<br />

behaviour. For example, adults seemed to have more difficulty in laying eggs (twisted<br />

ovipositor) on needles <strong>of</strong> fertilized trees than on unfertilized ones (Smirn<strong>of</strong>f & Bernier<br />

1973, Smirn<strong>of</strong>f & Valero 1975).<br />

Large-scale transplantations <strong>of</strong> red wood ants over long distances have only been tried a<br />

few times on the North American continent, and the costs involved can only be roughly<br />

estimated. The transplantations from central Europe were made co-operatively with<br />

scientific institutions (University <strong>of</strong> Pavia in Pavia, Italy, and CIBC in Delemont,<br />

Switzerland) and as a consequence had many unrevealed costs. The introduction from<br />

southern Manitoba, on the other hand, was funded completely by the Laurentian Forest<br />

Research Centre. Because the operations were experimental and <strong>of</strong> a scientific nature,<br />

no charges were made for the ant colonies collected. However, future large-scale<br />

operations may have to take this factor into consideration.<br />

The total costs can be divided among four operations: collecting, packaging, transportation,<br />

and release. The relative importance <strong>of</strong> each can vary considerably according<br />

to time involved, distances to be covered and accessibility <strong>of</strong> sites. Table 90 gives<br />

estimates <strong>of</strong> two large transplantations made to the province <strong>of</strong> Quebec: one from<br />

Manitoba in 1972, the other from Italy in 1973.<br />

As there has not been any measurable N. swainei popUlation in recent years in jack<br />

pine stands chosen for the introduction <strong>of</strong> ants, it has not yet been possible to measure<br />

Cost estimates for the transplantation <strong>of</strong> red wood ant colonies<br />

Collecting Dispersal Approximate<br />

Date (man days) Transportation (man days) Packaging total costs<br />

1972 25 Southern Manitoba to Quebec. 10 $1 ()()() $6500<br />

with motor vehicle and trailer<br />

- $2 ()()()<br />

1973 40 Northern Italy to Quebec 10 $1 ()()() $9500<br />

motor vehicles and by air<br />

- $5 ()()()


348 R. J. Finnegan and W. A. Smirnorr<br />

Smim<strong>of</strong>f, W.A.; Valero. J. (1975) Effets II moyen tenne de la fertilisation par urCe ou par potassium sur P;n/IS banks;ana L. el Ie<br />

comportement de ses insectcs devastateurs: tels que Neodiprion swa;nei (Hymenoptera: Tenthredinidae)<br />

et TOllmcye//a IlIImismalicum (Homoptcra: Coccidac). Calladian lO/lmal Forest Research 5,236-244.<br />

Smirn<strong>of</strong>f, W.A.; Feltes, J.J.; Haliburton, W. (1962) A virus disease <strong>of</strong> Swaine's jack pine sawfly. Neodiprion swaine; Midd. sprayed from an<br />

aircraft. Canadiall Entomologist 94,4TI-486.<br />

Tostowaryk, W. (1971a) Life history and behavior <strong>of</strong> PodisllS modUlUS (Hemiptera: Pentatomidae) in boreal forcst in Quebec. CQlladian<br />

Entomologist 103,662-679.<br />

Tostowaryk, W. (1971b) Relationship between parasitism and predation <strong>of</strong> diprionid sawflies. Annals <strong>of</strong> the Elltom%gica/ <strong>Society</strong> <strong>of</strong><br />

America 64,1424-1427.<br />

Tostowaryk, W. (1971c) The crfect <strong>of</strong> prey defense on the functional response <strong>of</strong> Podisus modutus (Hemiptera: Pentatomidae) to densities<br />

<strong>of</strong> the sawflies. Neodiprioll swainti and N. prall; banksiallQ (Hymenoptera: Neodiprionidae). Canadian<br />

Entomologist 104,61-69.<br />

Tostowaryk, W. (1972) Coleopterous predators <strong>of</strong> the Swaine jack pine sawDy Ntodiprion swainei Middleton (Hymenoptera: Diprionidae).<br />

Canadian Journal <strong>of</strong> Zoology 50,1139-1146.<br />

Tostowaryk, W. (1973) Population estimation and feeding behaviour <strong>of</strong> adult carnbids in jack pine stands. Quebec. CallQdian Forestry Service<br />

InfomuJtion Report Q·X·32. 23 pp.


Pest Status<br />

Chapter 60<br />

Operophtera bruceata (Hulst), Bruce<br />

Spanworm (Lepidoptera: Geometridae)<br />

W.G.H.IVES<br />

The Bruce spanworm, Operophlera bruceala (Hulst), attacks a variety <strong>of</strong> deciduous<br />

trees (Prentice 1963), and is very similar in appearance to the winter moth, Operophlera<br />

brumala (L.). Adults <strong>of</strong> the two species can be distinguished by differences in genitalia<br />

in males and in vestigial wing length in females (Eidt el al. 1966). Pupae can be separated<br />

by differences in the form <strong>of</strong> the cremaster, and larvae can be distinguished by the<br />

placement <strong>of</strong> the ocelli (Eidt & Embree 1968). In eastern <strong>Canada</strong>, sugar maples, Acer<br />

saccharum Marsh., and beech trees, Fagus grandi/olia Ehrh., are favoured hosts, but in<br />

western <strong>Canada</strong>, trembling aspen. Populus Iremuloides Michx., seems to be the principal<br />

host, particularly in Alberta. Although the Bruce spanworrn is not usually considered a<br />

major pest, a number <strong>of</strong> localized infestations, as well as more extensive outbreaks in<br />

which damage has been primarily due to this insect, have been reported by the Forest<br />

Insect and Disease Survey (Anon. 1952-79). The following account <strong>of</strong> infestation<br />

history is based on these reports, except where otherwise noted.<br />

In central and eastern <strong>Canada</strong>. the first reported damage occurred in a small stand <strong>of</strong><br />

sugar maple near Merrivale, Ontario, in 1951 and 1952. In Newfoundland a small<br />

infestation was reported on white birch, Belula papyri/era Marsh., in 1956. The first<br />

reported damage in the maritimes occurred in 1962, when a patch <strong>of</strong> sugar maple in<br />

Nova Scotia was severely defoliated. and a larger area <strong>of</strong> defoliation was reported in<br />

New Brunswick. Infestations in both Maritime Provinces increased in scope and intensity<br />

in 1963, and the outbreak reached its peak in 1964, when over 200 000 ha <strong>of</strong> severe<br />

defoliation occurred in Nova Scotia alone (Harrington 1968). Declines in populations<br />

were noted in 1965, and by 1966 the outbreak had collapsed. A short-lived outbreak also<br />

occurred in Quebec, where the insect reached epidemic proportions in many areas in<br />

1963. However, populations were lower in 1964, and only small pockets <strong>of</strong> defoliation<br />

were reported in 1965. At its peak, this outbreak covered some 39 000 km 2 (Martineau &<br />

Monnier 1966). In Ontario, where infestations <strong>of</strong> Bruce spanworrn were first reported,<br />

no further damage occurred until 1963, when a small pocket <strong>of</strong> aspen defoliation was<br />

reported near Sudbury. In 1965, three widely separated areas <strong>of</strong> severe sugar maple<br />

defoliation occurred: one was in Algonquin Park; another was on an island in Lake<br />

Huron; and the third was on hilltops and ridges near Sault Stet Marie. The outbreak<br />

decreased in intensity in 1966 and collapsed completely in 1967.<br />

In western <strong>Canada</strong>, damage to trembling aspen attributable to the Bruce spanworm<br />

was first reported from the Obed area <strong>of</strong> Alberta in 1956. Widespread areas <strong>of</strong> defoliation.<br />

much <strong>of</strong> it severe, were reported in 1957 and again in 1958, when about 130000 kml<br />

were moderately or heavily infested (Brown 1962). In 1959 some infestations had<br />

subsided, and all outbreaks collapsed in 1960. Severe defoliation <strong>of</strong> trembling aspen in<br />

northern British Columbia was reported in 1958 and 1959, but these infestations also<br />

collapsed in 1960.<br />

No further significant damage attributable to the Bruce spanworm was reported until<br />

1968, when extensive areas <strong>of</strong> moderate to severe defoliation <strong>of</strong> trembling aspen were<br />

again observed in Alberta. Both the intensity and size <strong>of</strong> this outbreak increased in 1969.<br />

Extensive damage also occurred in 1970, but only patches <strong>of</strong> aspen were severely<br />

defoliated in 1971. In Quebec, several maple groves were infested in 1970, and these<br />

were completely stripped in 1971 and again in 1972. The insect was somewhat less<br />

349


350 W. G. H. Ivcs<br />

Background<br />

Field Trials<br />

abundant in 1973. and populations continued to decline in 1974 and 1975. In Ontario,<br />

pockets <strong>of</strong> severe defoliation <strong>of</strong> sugar maple, white birch, and beech were reported along<br />

the south boundary <strong>of</strong> Algonquin Park in 1974, and the outbreak expanded and intensified in<br />

1975, apparently collapsing in 1976. Severe defoliation <strong>of</strong> trembling aspen also occurred<br />

in the Lake Nipigon area in 1976. The outbreak increased in area in 1977, but populations<br />

declined in most areas in 1978. In Alberta, severe defoliation by Bruce spanworm was<br />

again noted in the Obed area in 1978, and this outbreak continued in 1979 and 1980.<br />

Most <strong>of</strong> the reported outbreaks <strong>of</strong> Bruce spanworm have been <strong>of</strong> short duration.<br />

Consequently there was little if any tree mortality, even when defoliation was complete.<br />

One <strong>of</strong> the reasons for the short duration <strong>of</strong> the outbreaks, at least in eastern <strong>Canada</strong>,<br />

appears to be a nuclear polyhedrosis virus (NPV) which was described by Smirn<strong>of</strong>f<br />

(1964). The virus apparently causes appreciable mortality and thus hastens the collapse<br />

<strong>of</strong> the outbreaks. Although there were several reports <strong>of</strong> the virus occurring in the<br />

maritimes and in Quebec, none <strong>of</strong> the Forest Insect and Disease Survey reports from<br />

western <strong>Canada</strong> mentioned virus, and no experimental assessments <strong>of</strong> its virulence<br />

could be found in the literature. A small-scale field experiment, using NPV from eastern<br />

<strong>Canada</strong>, was therefore conducted in the Obed area <strong>of</strong> Alberta in 1979.<br />

A small supply <strong>of</strong> NPV was obtained by infecting larvae reared in a laboratory<br />

(J.c. Cunningham 1979 personal communications). Bruce spanworm eggs were obtained<br />

by collecting moss from around the base <strong>of</strong> infested trees near Obed in the fall <strong>of</strong> 1978.<br />

The eggs were incubated during the winter, and the emerging larvae were reared on an<br />

artificial diet in plastic cups. A total <strong>of</strong> 5 000 fifth-instar larvae was infected with virus.<br />

The dead larvae were collected and the virus purified by differential centrifugation; 4 x<br />

1011 polyhedral inclusion bodies (PIB) were obtained. The results <strong>of</strong> the field trial have<br />

already been described (Ives & Cunningham 1980) and only a brief description <strong>of</strong> the<br />

experiment need be given here. A strip <strong>of</strong> moderately infested trembling aspen, mostly<br />

under 10 m in height, was selected for the virus trials. Various concentrations <strong>of</strong> virus in<br />

an aqueous formulation containing 50 gil Shade® (Sandoz Inc.), 1 % Chevron® sticker<br />

and 25% v/v animal feed-grade molasses were applied until liquid dripped from the<br />

foliage (about 350 lIha) with a backpack mist blower. Virus concentrations <strong>of</strong> 10' and 10'<br />

PIB/ml were applied on 24 May. The larvae were in the first instar, the temperature was<br />

19°C, relative humidity 42%, and wind speed 0-8 kmlh. Leaves on the trees receiving<br />

10' PIB/ml were 1.5-2.0 em in diameter. The remaining trees had not yet flushed. A<br />

second application <strong>of</strong> virus, on a different group <strong>of</strong> trees, was made on 1 June. Larvae<br />

were in the second and third instars, the temperature was 21°C, R.H. 29%, and wind<br />

speed 3-11 kmlh. The leaves were fully flushed, and three concentrations <strong>of</strong> virus (IO',<br />

10', and 10' PIB/ml) were tested.<br />

Two methods <strong>of</strong> assessing impact were used: determination <strong>of</strong> percentage mortality<br />

by rearing <strong>of</strong> individual larvae until death or pupation, and estimation <strong>of</strong> percentage<br />

defoliation by examining large numbers <strong>of</strong> individual leaves to determine the approximate<br />

amount eaten. Neither method was completely satisfactory. Aseptic techniques were<br />

not used in the collection <strong>of</strong> larvae, and all larvae for each sample were placed in the same<br />

container, so that contamination after collection probably occurred. The foliage samples<br />

were not collected until mid-August, and there was some evidence that the petioles <strong>of</strong><br />

the more severely damaged leaves had fallen by then. Consequently, the larval rearings<br />

may have tended to overestimate rates <strong>of</strong> infection, while the defoliation may have been<br />

underestimated. A further complication was the presence <strong>of</strong> naturally occurring virus.


Evaluation <strong>of</strong> Control Attempt<br />

Recommendations<br />

Literature Cited<br />

0l'('rol'illt'rtl bruceattl (Hulst). 351<br />

Two samples <strong>of</strong> late-instar larvae were collected about 0.4 km from the experimental<br />

area. Microscopic examination <strong>of</strong> living larvae from those samples revealed that 13 and<br />

2t % respectively were infected with NPV (Ives & Cunningham 1980).<br />

The percentages <strong>of</strong> virus-caused mortality <strong>of</strong> larvae collected 7 days after the 24 May<br />

application were 80 and 98% for larvae receiving 10' and 10' PIB/ml, compared to a<br />

mean <strong>of</strong> 52% for the two untreated check areas. Corresponding mortalities among larvae<br />

collected 10 days after the 1 June application were 94. 100. and 100% for dosages <strong>of</strong> 10'.<br />

to' and to" PIB/ml respectively, compared to a mean mortality <strong>of</strong> 68% for larvae<br />

collected in the check area. All treatments afforded some degree <strong>of</strong> foliage protection,<br />

but only the 24 May application <strong>of</strong> 10' PIB/ml and the 1 June application <strong>of</strong> to' PIB/ml<br />

provided protection that could be detected visually.<br />

Bruce spanworm NPV appears to be virulent, on the basis <strong>of</strong> the limited amount <strong>of</strong><br />

information available. However, it is expensive to produce, because it must be propagated<br />

in a small host. Its use in artificially regulating populations <strong>of</strong> Bruce spanworm does not<br />

currently seem to be practical, partly because <strong>of</strong> cost and partly because the rate <strong>of</strong><br />

natural virus infection usually appears to terminate outbreaks before appreciable damage<br />

occurs.<br />

Ultimately, this NPV might have potential as an applied biological insecticide, but this<br />

will not occur unless more economical methods for propagating insect viruses are<br />

developed. No further testing <strong>of</strong> this virus is recommended in the immediate future.<br />

Anonymous (1952 - 1979) Annual reports <strong>of</strong> the Forest Insect and Disease Survey. 1951-1976. Ottawa. Ontario; Canadian Forestry Senice.<br />

Brown. C.E. (1962) The life history and dispersal <strong>of</strong> the Bruce spanworm. Operophtera bruceala (Hulst) (Lepidoptera: Geometridae).<br />

Canadian Entomologist 94.1103-1107.<br />

Eidt. D.C.; Embree, D.G. (1968) Distinguishing larvae and pupae <strong>of</strong> the winter moth. Operophtera brumata. and the Bruce spanworm. O.<br />

bruceata (Lepidoptera: Geometridae). Canadian Entomologist 100.536-539.<br />

Eidt, D.C.; Embree. D.G.; Smith. C.C. (1966) Distinguishing adults <strong>of</strong>the winter moth, Operophtera brumata (L.). and Bruce spanworm. O.<br />

brumata (Hulst) (Lepidoptera: Geomelridae). Canadian Entomologist 98.258-261.<br />

Harrington. W. (1968) Relation <strong>of</strong> light trap catches <strong>of</strong> Bruce spanworm to infestations in ccntral Nova Seotia. Canadian Department <strong>of</strong><br />

Fisheries and Forestry Bi-monthly Research Notes 24(5),39-40.<br />

I vcs, W. G. H. ; Cunningham, J. C. (1980) Application <strong>of</strong> nuclear polyhedrosis virus to control Bruce spanworm (Lepidoptera: Geometridae).<br />

Canadian Entomologist 112,741-744.<br />

Martineau. R.; Monnier. C. (1966) Recent outbreak <strong>of</strong> the Bruce spanworm in Quebec. Canadian Department <strong>of</strong> Forestry and Rural<br />

Development Bi-monthly Research Notes 22(6).5.<br />

Prentice. R.M. (1963) Forest Lepidoptera <strong>of</strong> <strong>Canada</strong> recorded by the Forest Insect Survey. Canadian Department <strong>of</strong> Forest Entomology and<br />

Pathology Branch, Ottawa, Publication 1013.<br />

Smirn<strong>of</strong>f. W.A. (1964) A nuc\eopolyhedrosis virus <strong>of</strong> Operophtera bruceata (Hulst) (Lepidoptera: Geometridae). Journal <strong>of</strong> Insect<br />

Pathology 6.384-386.


Blank Page<br />

352


Pest Status<br />

Background<br />

Chapter 61<br />

Operophtera brumata (L.), Winter Moth<br />

(Lepidoptera: Geometridae)<br />

D.O. EMBREE and I.S. OTVOS<br />

The winter moth. Operophtera brumata (L.). was accidentally introduced into Nova<br />

Scotia. probably in the mid 1930s. and later spread to New Brunswick and Prince<br />

Edward Island. Recently it has been found in British Columbia. and in western Oregon<br />

in the United States.<br />

Populations <strong>of</strong> the insect in the maritimes have remained consistently low during the<br />

20-year review period because <strong>of</strong> the establishment and population build-up <strong>of</strong> the<br />

tachinid parasitoid Cyzenis albicans (Fall.) and the ichneumonid parasitoid<br />

Agrypon flaveolatum (Gravely) in 1961 (Embree 1971a). The gradual spread <strong>of</strong> the<br />

insect since it was first detected in 1949 has virtually ceased and it is almost non-existent<br />

in forested areas. However. populations have persisted in apple. Malus spp .• orchards<br />

and on shade trees. particularly oaks. Quercus spp .• and lindens. TWa spp .• (Sterner &<br />

Davidson 1981). Chemical control is occasionally required in orchards but generally<br />

parasitoid numbers have remained sufficiently high to keep the winter moth under<br />

control.<br />

In British Columbia. the winter moth is found in and around Victoria on Vancouver<br />

Island and in Richmond near Vancouver. As in Nova Scotia (Hawboldt & Cuming 1950),<br />

the insect was initially misidentified and was not suspected to be O. brumata until 1976.<br />

4 years after it was causing significant damage to Garry oak. Quercus garryana Doug!..<br />

in Victoria. By 1977 the outbreak had expanded over large areas and an estimated 120<br />

kml were severely affected in Victoria on the Saanich Peninsula. The centre <strong>of</strong> the<br />

outbreak appeared to be in the Wilkinson Road area in Victoria where several commercial<br />

nurseries are located. suggesting that the insect might have been introduced into British<br />

Columbia on nursery stock (Gillespie & Finlayson 1981). In British Columbia the larvae<br />

were found to feed on Acer. Crataegus. Malus, Prunus, Populus, Quercus. and Salix<br />

spp .• and other deciduous trees and shrubs and to pose a threat to commercial orchards<br />

and shade trees in urban areas. Parasitism by native species <strong>of</strong> parasitoids is less than 5%<br />

(Gillespie et al. 1978).<br />

In the United States. outbreaks <strong>of</strong> this pest are confined to Multnomach, Clackamus.<br />

and Washington counties in western Oregon. where the insect is found in abandoned<br />

orchards and is becoming a pest <strong>of</strong> filberts. Corylus spp.<br />

Biological control <strong>of</strong> the winter moth in the maritimes by the introduced parasitoids C.<br />

albicans and A. flaveolatum has been successful and remains so because it is helped by<br />

the behaviour <strong>of</strong> the winter moth which results in most <strong>of</strong> the host larvae starving before<br />

suitable food is available. Most <strong>of</strong> the time. winter moth hatching occurs before budburst <strong>of</strong><br />

its principal forest host. red oak. Quercus rubra L.. and other forest species. because less<br />

heat is required for hatching than for budburst. Over 98% <strong>of</strong> the newly hatched larvae<br />

may starve because they cannot feed on closed buds and the few survivors are then<br />

controlled by the parasitoids (Embree 1965).<br />

On hatching. larvae initially spin downwards and are then carried up to the tree<br />

foliage on rising air currents. Such currents are most apt to occur in the early morning<br />

when the air begins to warm and rise. By mid-morning. when the land is fully heated.<br />

353


354 D. G. Embree and I. S. Olvos<br />

Releases and Recoveries<br />

Agrypon fla.-eo/atum<br />

(Grav.)<br />

(Hymenoptera:<br />

Ichneumonidae)<br />

Cyzenis a1bicans<br />

(FaIl.) (Dlptera:<br />

Tachinidae)<br />

wind patterns develop. Winter moth hatching is synchronized to this pattern <strong>of</strong> wind<br />

movement, with hatching beginning at dawn and virtually ceasing by mid-morning. By<br />

this efficient means larvae are transported from hatching sites in crevices and lichens on<br />

the lower tree trunks to the tree crowns. Once there, the larvae either survive or die<br />

depending on whether the buds have opened. The wind may also transport the larvae<br />

several miles, but once established on the foliage, they do not dislodge easily until the<br />

late fifth instar in early summer, when they drop to the ground to pupate (Embree 1970).<br />

No other means <strong>of</strong> dispersal exists because the females, which are active in November<br />

and December, have vestigial wings and are flightless.<br />

No releases <strong>of</strong> parasitoids have been made in the maritimes since 1965 (Embree 1971b).<br />

Releases <strong>of</strong> parasitoids in British Columbia were conducted in 1979 and 1980 and are<br />

listed in Table 91.<br />

The British Columbia operation was financed and administered by the Crop Protection<br />

Branch (formerly Entomology and Plant Pathology Branch) <strong>of</strong> the British Columbia<br />

Ministry <strong>of</strong> Agriculture and Food in co-operation with the Pacific and Maritimes Forest<br />

Research Centres <strong>of</strong> the Canadian Forestry Service. Collections <strong>of</strong> last instar larvae<br />

were made in apple orchards in Nova Scotia in 1978 and 1979. Collections were also<br />

made in Germany by the Commonwealth Institute <strong>of</strong> Biological Control.<br />

Winter moth pupae were sent to the Biocontrol Unit <strong>of</strong> the Research Branch <strong>of</strong><br />

Agriculture <strong>Canada</strong> in Ottawa, where pupae were cold-treated and then reared under<br />

quarantine conditions. Only the adult parasitoids were shipped to British Columbia from<br />

the German collection but about two-thirds <strong>of</strong> the Nova Scotia collection were sent to<br />

Victoria for parasitoid rearing. The remainder were reared in Ottawa and the adults <strong>of</strong><br />

the two parasitoids were sent to Victoria.<br />

Based on studies in Nova Scotia, the control strategy in British Columbia has been to<br />

treat the infested area with repeated releases, each consisting <strong>of</strong> a minimum <strong>of</strong> 50 mated<br />

females <strong>of</strong> each species.<br />

This parasitoid beromes decreasingly effective as host densities increase, and is therefore most<br />

effective at low population density (Embree 1966). The adult parasitoids develop in<br />

winter moth pupae and emerge the following spring to parasitize early instars <strong>of</strong> host<br />

larvae.<br />

The first introduction <strong>of</strong> A. Jlaveolatum was in 1979 when a total <strong>of</strong> 1654 males and<br />

1 700 females was released at 33 locations in the greater Victoria area in stands infested<br />

with winter moth. Fifty pairs <strong>of</strong> parasitoids were released in each location after being<br />

held in mating cages for 1-3 days. The parasitoids in the cages were provided with twigs<br />

<strong>of</strong> cherry blossoms and bottles <strong>of</strong> sugared water with feeding wicks. In 1980 an additional<br />

2213 males and 2 813 females were released at 25 locations; <strong>of</strong> these a small proportion<br />

came from Germany, and the rest from Nova Scotia. This parasitoid has not yet been<br />

recovered in British Columbia.<br />

This parasitoid exhibits a sigmoid functional response curve and is most effective at high<br />

host densities (Embree 1966). The female, which can lay as many as 1300 eggs, oviposits<br />

in the vicinity <strong>of</strong> feeding damage by defoliators. Associated species such as the fall<br />

cankerworm, Alsophila pometaria (Harr.), which are resistant to parasitism by C.<br />

albicans, also consume the parasitoid eggs. With increasing winter moth populations C.<br />

albicans becomes increasingly effective.


Blank Page<br />

358


Pest Status<br />

Background<br />

Chapter 62<br />

Orgyia leucostigma (J.E. Smith),<br />

Whitemarked Tussock Moth (Lepidoptera:<br />

Lymantriidae)<br />

D.G. EMBREE, D.E. ELGEE and G.F. ESTABROOKS<br />

The whitemarked tussock moth, Orgyia leucostigma (J.E. Smith), is a noxious insect in<br />

eastern <strong>Canada</strong>, particularly in Nova Scotia. It is primarily a forest pest, capable <strong>of</strong><br />

killing entire stands <strong>of</strong> mature balsam fir, Abies balsamea (L.) Mill., in a year and <strong>of</strong><br />

defoliating extensive areas <strong>of</strong> hardwoods. It is particularly devastating to balsam fir<br />

Christmas trees which it deforms, defoliates, or kills. The white marked tussock moth is<br />

also a pest in urban areas where it feeds on a wide variety <strong>of</strong> hardwood shade trees,<br />

crawls over houses and lawns, and spins unsightly cocoons on trees and shrubs. Some<br />

people are allergic to the larval setae. Moreover, O. leucostigma is a pest <strong>of</strong> blueberries<br />

and is the target <strong>of</strong> an extensive spray programme to protect the commercial crop.<br />

Egg masses are laid on empty female cocoons and covered with froth. On hatching,<br />

larvae feed gregariously on the froth. Once this is consumed, larvae spin downwards on<br />

silken threads and are carried great distances by the wind. On occasion, during severe<br />

outbreaks, swarms <strong>of</strong> airborne larvae appear as a grey mist and the insects can cover the<br />

surface <strong>of</strong> small lakes and ponds when they land.<br />

The larvae feed on a wide variety <strong>of</strong> hardwoods and conifers such as tamarack, Larix<br />

laricina (Du Roi) K. Koch, but their primary host is balsam fir. Larvae feed in the shade<br />

on the underside <strong>of</strong> leaves and twigs. This feeding habit causes a unique form <strong>of</strong> damage<br />

to balsam fir Christmas trees because larvae also feed on the thin bark <strong>of</strong> developing<br />

leaders and upper lateral branches causing them to warp downwards. The result is a<br />

pronounced crook in the stems <strong>of</strong> surviving trees, visible for years afterwards. In heavy<br />

infestations, trees are festooned with silk.<br />

Pupation usually occurs on the underside <strong>of</strong> twigs and branches or on the stems <strong>of</strong> host<br />

trees, but it can occur almost anywhere, the yellowish white cocoons <strong>of</strong>ten being spun<br />

on fences, houses, or logs. Females remain on the surface <strong>of</strong> their cocoons when they<br />

emerge, mate with winged males, and lay from 100 to over 500 eggs.<br />

Outbreaks are controlled ultimately by a nuclear polyhedrosis virus (NPV). These outbreaks<br />

<strong>of</strong> O. leucostigma, at least in Nova Scotia, appear to originate in balsam fir<br />

stands at high elevations and occur at intervals <strong>of</strong> 6-10 years between outbreaks. The<br />

infestation spreads by wind-borne larvae and eventually large areas <strong>of</strong> balsam fir and<br />

hardwoods are defoliated. Such outbreaks become widespread and last for 5-6 years,<br />

then collapse because <strong>of</strong> the gradual spread <strong>of</strong> the virus disease. The initial infestation<br />

collapses in about 3 years, subsequent infestations in 2-3 years, and in the final stages <strong>of</strong><br />

the outbreak infestations appear and collapse in the same year.<br />

Over 25 species <strong>of</strong> parasitoids attack the whitemarked tussock moth but the only<br />

control they have over populations is to kill survivors following the collapse <strong>of</strong> outbreaks<br />

caused by NPV. As a result, whitemarked tussock moth populations are reduced suddenly<br />

to extremely low levels, which delays the build-up <strong>of</strong> subsequent outbreaks.<br />

359


360 D. G. Embree. D. E. Elgee and G. F. Estabrooks<br />

Field Trials<br />

Evaluation <strong>of</strong> Control Attempts<br />

No attempt has been made to control this insect through the introduction <strong>of</strong> additional<br />

parasitoid species. The emphasis has been on establishing control through the manipulation<br />

<strong>of</strong> the virus (Cunningham 1972, Elgee 1975). Virus suspensions can be used as<br />

insecticides and have been tested successfully on small trees.<br />

A major attempt at biological control was made in 1975 to initiate an epizootic in a large<br />

population <strong>of</strong> the whitemarked tussock moth in the hope <strong>of</strong> shortening the length <strong>of</strong> an<br />

outbreak; it has been reported by Kurstak (in press). During the previous year the<br />

suspected start <strong>of</strong> a general outbreak <strong>of</strong> the whitemarked tussock moth was detected in an<br />

IS-ha stand <strong>of</strong> balsam fir near Castlereagh, Colchester County, Nova Scotia. A virus<br />

suspension was prepared from third-instar larvae reared from eggs collected at the site<br />

and infected with the Nova Scotia strain <strong>of</strong> NPV: the virus had been saved and stored<br />

from the last outbreak in 1965.<br />

On the evening <strong>of</strong> 9 July 1975, 204 I <strong>of</strong> an NPV suspension containing IO.S x HY'<br />

polyhedral inclusion bodies (PIB) per millilitre mixed with 204 I <strong>of</strong> a virus spray adjuvant<br />

manufactured by Sandoz Inc. and a fluorescent dye were sprayed from 30 m on IS ha <strong>of</strong><br />

mature balsam fir with a Cessna Ag-truck aircraft equipped with a boom and nozzle<br />

sprayer. Whitemarked tussock moth larvae were mainly first instars with some early<br />

second instars.<br />

There were no true controls (untreated areas) but larvae collected before the spray<br />

application and reared were free from disease. Twenty-four sample trees were established<br />

in a line bisecting the spray block. However, one edge <strong>of</strong> the block was not sprayed so<br />

that trees 20-24 were outside the block. No disease was detected in larvae collected on<br />

tree 24 and only light mortality (33%) occurred on trees 20-23. Mortality <strong>of</strong> larvae<br />

collected 1 day after spraying from trees 1-19, and reared on foliage from the sprayed<br />

area, averaged 84%, whereas that <strong>of</strong> survivors collected 2 weeks later was 64%. Field<br />

populations declined by 95%, as compared to 67% on the unsprayed trees, and defoliation<br />

was light. From the air, the outline <strong>of</strong> the sprayed area was clearly discernible in a<br />

severe outbreak that developed around the initial Castlereagh infestation. No appreciable<br />

population reappeared in the sprayed area. The outbreak in the surrounding areas intensified<br />

in 1976, covering a radius <strong>of</strong> 16 km, but collapsed from the virus disease in all but<br />

the periphery during the same year.<br />

Early in 1976, virus suspensions were dispersed at two locations along the periphery<br />

<strong>of</strong> the 1975 outbreak using an explosive device. Canisters containing 1.9 I <strong>of</strong> suspension<br />

were fired into the air using a home-made mortar and exploded over the forest canopy.<br />

Virus particles were dispersed over an area <strong>of</strong> 1000-2000 m l depending on wind velocity.<br />

However, by this time the virus was already present in the population. Eventually the<br />

outbreak spread throughout the province, collapsing everywhere by 1975.<br />

The experiment showed that the virus can be used as an effective biological insecticide.<br />

but whether it can be used to initiate a widespread epizootic is uncertain. The rapid<br />

spread <strong>of</strong> the O. ieucosligma infestation in 1975 resulted from dispersion <strong>of</strong> newly hatched<br />

larvae, most <strong>of</strong> which occurred before spraying took place. Moreover. low populations<br />

<strong>of</strong> larvae existed in the area immediately surrounding the initial 18 ha <strong>of</strong> detected outbreak.<br />

The outbreak in 1976 was widespread and extremely intense; even though it collapsed<br />

just 2 years after it appeared, it is unlikely that the source <strong>of</strong> the epizootic was the IS-ha<br />

spray block. Had the initial outbreak been detected in 1973 and sprayed in 1974 a clearer


Recommendations<br />

Literature Cited<br />

Orgyilllelicosligll/(/ (J. E. Smith), 361<br />

understanding <strong>of</strong> the role <strong>of</strong> the virus in whitemarked tussock moth epidemiology might<br />

have been gained.<br />

The pattern <strong>of</strong> development <strong>of</strong> whitemarked tussock moth outbreaks presents an<br />

opportunity for an elegant approach to biological control; but there are major difficulties.<br />

The most promising strategy is to artificially accelerate the normal virus epizootic that<br />

has occurred consistently in every past outbreak <strong>of</strong> this insect. The obvious advantage<br />

<strong>of</strong> such a strategy is that the virus is used as an inoculum rather than as a blanket<br />

insecticide. The resulting epizootic will be one that inevitably would have occurred<br />

later, a consideration that might ease the registration approval <strong>of</strong> this NPV by Canadian<br />

authorities.<br />

Historically each whitemarked tussock moth outbreak in Nova Scotia appears to have<br />

begun in one or two small locations (epicentres). Virus material from the 1974-79 outbreak<br />

has been stored and the techniques <strong>of</strong> introducing the virus into a developing<br />

outbreak have been shown to be workable. However, two major problems must be<br />

addressed: first, early detection <strong>of</strong> epicentres requires constant monitoring <strong>of</strong> likely<br />

sites for a period <strong>of</strong> up to 10 years; and second, whitemarked tussock moth hatching<br />

occurs over a period <strong>of</strong> at least 2 weeks, so although larvae feed on the froth covering <strong>of</strong><br />

egg masses for 2 or 3 days, many larvae are likely to disperse before the virus is applied<br />

during peak egg hatching.<br />

A second strategy might be to employ sex pheromones that could be used to disrupt<br />

mating in epicentres. Success would depend on the almost immediate detection <strong>of</strong> these<br />

epicentres before they become large enough to develop a reservoir <strong>of</strong> dispersing larvae.<br />

Cunningham, J.e. (1972) Preliminary studies <strong>of</strong> nuclear polyhedrosis viruses infecting the whitemarked tussock moth, Orgyia lel/costigma.<br />

Canadian Forestry Service In.tect Pathology Research Institl/te, Sal/It Ste. Marie, Ontario, Information<br />

Report.<br />

Elgee, D.E. (1975) Persistence <strong>of</strong> a virus <strong>of</strong> the whitemarked tussock moth on balsam fir foliage. Canadian Forestr}' Service Bi-monthly<br />

Research Notes 31(5),33-34.<br />

Kurstak, E. (in press) Microbial and viral pesticides. New York; Marcel Dekker.


Blank Page<br />

362


Pest Status<br />

Background<br />

Field Trials<br />

Nuclear polyhedrosis<br />

virus<br />

Chapter 63<br />

Orgyia pseudotsugata (McDunnough),<br />

Douglas-fir Tussock Moth (Lepidoptera:<br />

Lymantriidae)<br />

J.C. CUNNINGHAM and R.F. SHEPHERD<br />

The Douglas-fir tussock moth, Orgyia pseudotsugata (McDunn.), occurs in the semiarid<br />

interior <strong>of</strong> British Columbia and in parts <strong>of</strong> Washington, Idaho, California, Nevada,<br />

Colorado, Arizona, and New Mexico in the United States. In <strong>Canada</strong>, Douglas fir,<br />

Pseudotsuga menziesii (Mirb.) Franco, is the preferred host; but Engelmann spruce,<br />

Picea engelmannii Parry, is also attacked and later instar larvae will feed on ponderosa<br />

pine, Pinus ponderosa Laws., when forced to abandon totally defoliated Douglas fir.<br />

Small larvae eat the underside <strong>of</strong> new needles. The later instar larvae may eat entire<br />

older needles or sever them near the base and leave them entwined in silk webbing.<br />

Heavily infested trees have a reddish-brown appearance. Trees may die after 1 year <strong>of</strong><br />

defoliation, but mortality occurs more commonly after 2 or more years <strong>of</strong> severe<br />

defoliation (Johnson & Ross 1967). Severely damaged trees are also susceptible to<br />

attack by the Douglas-fir beetle, Dendroctonus pseudotsugae Hopkins.<br />

Since 1916, when the first records were kept (Sugden 1957), six outbreaks <strong>of</strong> Douglasfir<br />

tussock moth occurred in British Columbia. The latest outbreak began in 1971 and<br />

collapsed in 1976; a further outbreak was predicted for 1981 (Sterner & Davidson 1981).<br />

Characteristically, high populations <strong>of</strong> Douglas-fir tussock moth appear in relatively<br />

small patches <strong>of</strong> forest <strong>of</strong> 1-5 ha. The infestation normally extends from these areas in<br />

subsequent years, but more striking is the appearance <strong>of</strong> new outbreaks <strong>of</strong>ten many<br />

kilometres from the original sites. After an average <strong>of</strong> 5 years (3-6) outbreaks collapse,<br />

the factors responsible being nuclear polyhedrosis virus (NPV), egg parasitoids, and<br />

starvation. However, following several years <strong>of</strong> severe defoliation, the trees frequently<br />

die before this population collapse occurs. Hence attempts to regulate the population<br />

with biological or chemical agents should be made early in the outbreak cycle.<br />

Two types <strong>of</strong> nuclear polyhedrosis virus (NPV) have been found in populations <strong>of</strong><br />

Douglas-fir tussock moth. In one type the rod-shaped virus particles are embedded<br />

singly in polyhedral inclusion bodies (PIB) and in the other they are embedded in<br />

bundles (Hughes & Addison 1970). These are referred to as single-embedded NPV<br />

(SNPV) and multiple-embedded NPV (MNPV). An SNPV isolated from the whitemarked<br />

tussock moth. O. leucostigma (J.E. Smith), is also pathogenic for Douglas-fir tussock<br />

moth larvae.<br />

Small-scale ground-spray trials in 1962 with field-collected NPV (possibly a mixture<br />

<strong>of</strong> SNPV and MNPV) gave encouraging results (Morris 1963). The first aerial spray trial<br />

in British Columbia was conducted in 1974 using Douglas-fir tussock moth MNPV<br />

363


366 J. C. Cunningham and R. F. Shepherd<br />

Evaluation <strong>of</strong> Control Attempts<br />

Recommendations<br />

Some <strong>of</strong> the treatments were designated operational and others experimental.<br />

Twenty-one days after spraying, larval mortality due to the B.t. treatments averaged<br />

34% with a high <strong>of</strong> 57% with Thuricide® and a low <strong>of</strong> 13% with Dipel®. A double<br />

application <strong>of</strong> both materials and a higher volume <strong>of</strong> Thuricide® gave about 20% more<br />

control than single applications or the lower volume. Increasing the dosage <strong>of</strong> Dipel®<br />

gave no significant change, but using molasses in the tank mix instead <strong>of</strong> sorbitol gave a<br />

significant increase in the effectiveness <strong>of</strong> Dipel® and brought it to the same level as<br />

Thuricide® .<br />

Neither single nor double applications <strong>of</strong> Dipel® provided adequate foliage protection-<br />

59% defoliation was recorded on untreated check plots and 50-58% on Dipel®-treated<br />

plots. Better foliage protection was recorded with Thuricide 81 with only 18-25% defoliation<br />

on treated trees.<br />

The various operational and experimental treatments with B.t. gave a wide range <strong>of</strong><br />

results. However, the best population reductions due to treatment were less than 60%.<br />

This did not provide adequate foliage protection and did not prevent the Douglas-fir<br />

tussock moth population increasing to high densities in the next generation.<br />

Results with B.t. on Douglas-fir tussock moth in 1975 were considered unsatisfactory<br />

and applications <strong>of</strong> B.t. on other defoliating lepidopterous forest pests during the last<br />

decade in British Columbia have generally proved disappointing. Recent improvements<br />

in application technology <strong>of</strong> B.t. have resulted in improved control <strong>of</strong> spruce budworm,<br />

Choristoneurafumiferana (Oem.), and similar improvements may eventually be possible<br />

with Douglas-fir tussock moth. On the other hand, results with NPV are already most<br />

encouraging and it appears that this biological control agent can now provide a useful<br />

tool for the regulation <strong>of</strong> Douglas-fir tussock moth populations.<br />

Douglas-fir tussock moth MNPV was registered by the Environmental Protection<br />

Agency in the United States in 1975 under the name TM Biocontrol-1. It is proposed to<br />

apply for Canadian registration in 1982. At present this virus is not commercially<br />

produced and this poses two major problems that need solving: a source <strong>of</strong> supply; and<br />

production at a reasonable cost. Small amounts, sufficient to treat about 400 ha annually,<br />

are produced at the Forest Pest Management Institute, Sault Ste. Marie, Ontario, and<br />

larger amounts are produced for use in the United States by staff <strong>of</strong> the U.S. Forest<br />

Service, Corvallis, Oregon. A figure <strong>of</strong> $40 US per hectare was quoted for production at<br />

Corvallis in late 1980 (M.E. Martignoni personal communication) and this figure is<br />

probably even higher for material from Sault Ste. Marie. This cost figure is based on a<br />

dosage <strong>of</strong> 250 x 10' PIBlha. This dosage can probably be reduced to 125 x 10 9 PIBlha<br />

(I1nytzky et al. 1977), and possibly even lower.<br />

The mountainous terrain and consistently low relative humidity encountered in the<br />

interior <strong>of</strong> British Columbia pose problems for aerial application <strong>of</strong> pest control agents<br />

and particularly for application <strong>of</strong> aqueous spray formulations. Use <strong>of</strong> an oil-based<br />

formulation should be investigated, for this may well enhance the deposit and facilitate<br />

application <strong>of</strong> lower dosages.<br />

Ideally, application <strong>of</strong> a virus initiates an epizootic in the pest insect population,<br />

regulating the pest either in the year <strong>of</strong> application if applied on early-instar larvae, or in<br />

the subsequent year. Some NPVs have the potential to initiate epizootics and others<br />

have not; the status <strong>of</strong> Douglas-fir tussock moth NPV has not been established and the


Literature Cited<br />

Orgyia psellClolsllgata (McDunnough). 367<br />

virus has been applied in the same manner as a chemical pesticide. Although adequate<br />

control has been achieved, the material is too scarce and too costly to be applied in this<br />

way. Alternative methods should be investigated, such as seeding the NPV into the<br />

insect population. This could be tested by spraying widely spaced swaths to give partial<br />

coverage <strong>of</strong> the infested forest. Spread <strong>of</strong> the disease would then be relied on to regulate<br />

the Douglas·fir tussock moth population.<br />

Among the few viruses currently being developed as biological control agents <strong>of</strong><br />

forest insect pests in <strong>Canada</strong>, Douglas.fir tussock moth NPVs have proved to be<br />

effective pest management tools. Further testing and an attempt to register MNPV in<br />

<strong>Canada</strong> for operational use are strongly advocated. Improved production methods,<br />

lower dosage rates, and different strategies for its use may reduce the cost <strong>of</strong> treatment<br />

and make MNPV more attractive to forest managers.<br />

Furtht!r efficacy trials with B.t. may be warranted, as better strains and improved<br />

application technology become available. This work, however, should have lower<br />

priority than the virus work described above.<br />

Abbott, W.S. (1925) A method <strong>of</strong> computing the effectiveness <strong>of</strong> an insecticide. Journal <strong>of</strong> Economic Entomology 18.265-267.<br />

Hughes, K.M.; Addison. R.B. (1970) Two nuclear polyhedrosis viruses <strong>of</strong> the Douglas·fir tussock moth. Journal <strong>of</strong> Invertebrate Pathology<br />

16,196-294.<br />

I1nytzky, S.; McPhee. J.R.; Cunningham. J.C. (19n) Comparison <strong>of</strong> field'propagated nuclear polyhedrosis virus from Douglas-fir tussock<br />

moth with laboratory-produced virus. Canadian Department <strong>of</strong> Fisherits and Forts"y Bi-monthly<br />

Research Notts 33(1).5-6.<br />

Johnson. P.C.; Ross. D.A. (1967) Douglas-fir tussock moth. Hemerocampa (Orgyia) pseudotsugata McDunnough. In: Davidson. A.G.;<br />

Prentice. R.M. (Camps. and Eds.) Important forest insects and diseases <strong>of</strong> mutual concern to <strong>Canada</strong>. the<br />

United States and Mexico. Onawa; Canadian Department <strong>of</strong> Forestry and Rural Development. pp.<br />

105-107.<br />

Morris. O.N. (1963) The natural and artificial control <strong>of</strong> the Douglas-fir tussock moth. Orgyia pseudotsugata McDunnough. by a nuclear<br />

polyhedrosis virus. Journal <strong>of</strong> Insect Pathology 5.401-414.<br />

Shepherd, R.F. (Ed.) (1980) Operational field trials against Douglas-fir tussock moth with chemical and biological insecticides. Canadian<br />

Forestry Service. Victoria. British Columbia. Information Report BC-X-201. 19 pp.<br />

Stelzer. M.; Neisess. J.; Cunningham. J.C.; McPhee. J.R. (1m) Field evaluation <strong>of</strong> baculovirus stocks against Douglas-fir tussock moth in<br />

British Columbia. Journal <strong>of</strong> Economic Entomology 70.243-246.<br />

Sterner. T.E.; Davidson. A .G. (camp.) (1981) Forest insect and disease conditions in <strong>Canada</strong> 1980. Canadian Fortslry Service Forest Insect and<br />

Disease Survey. 43 pp.<br />

Sugden, B.A. (1957) A brief history <strong>of</strong> the Douglas-fir tussock moth, Hemerocampa pseudotsugata McD., in British Columbia. Proceedings<br />

<strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> British Columbia 54.37-39.


Blank Page<br />

368


Pest Status<br />

Background<br />

Chapter 64<br />

Pristiphora erichsonii (Hartig), Larch<br />

Sawfly (Hymenoptera: Tenthredinidae)<br />

W.G.H. IVES and J.A. MULDREW<br />

The larch sawfly, Pristiphora erichsonii (Htg.), is a major pest <strong>of</strong> larches, LArix spp.,<br />

in North America. It is the principal defoliator <strong>of</strong> tamarack, L. laricina (Du Roi) K.<br />

Koch, and also attacks western larch, L. occidentalis Nutt., and alpine larch, L. Iyallii<br />

ParI. Plantings <strong>of</strong> the various European and Asian species and their hybrids are also<br />

vulnerable (Turnock & Muldrew 1971a). Outbreaks have been recorded in <strong>Canada</strong> since<br />

late in the nineteenth century. Early outbreaks have been discussed by McGugan &<br />

Coppel (1962), and Turnock & Muldrew (1971a) summarized depredations for the<br />

decade 1959-68. Conditions since then have been outlined in the annual reports <strong>of</strong> the<br />

Forest Insect and Disease Survey (Anon. 1970-79). The folIowing account <strong>of</strong> infestation<br />

history from 1969 to 1980 is based primarily on these reports. Areas in which infestations<br />

<strong>of</strong> the larch sawfly have caused moderate or severe defoliation during this period are<br />

shown in Fig. 18.<br />

In Newfoundland, tamarack stands in the western and central parts <strong>of</strong> the province<br />

were moderately or severely defoliated in 1975, but most infestations colIapsed in 1976.<br />

A resurgence occurred along the west coast in 1979. In Labrador, an outbreak developed in<br />

1975, and by 1978 some tree mortality had been reported. In the maritimes, remnants <strong>of</strong><br />

an earlier outbreak persisted into the period, particularly in southern New Brunswick<br />

and central Nova Scotia. In eastern Prince Edward Island, infestations increased in<br />

1970 and populations did not colI apse until 1976. By this time over 30% <strong>of</strong> the trees were<br />

dead, and many more had dead tops. In Nova Scotia, populations started increasing in<br />

1970 and by 1975 tamarack stands in most <strong>of</strong> the province were affected. Populations in<br />

the eastern half <strong>of</strong> the province colIapsed in 1976, but the decline was more gradual in the<br />

western half and some infestations still persisted in 1980. In New Brunswick, larch<br />

sawfly damage during the period 1969-80 was confined to a number <strong>of</strong> relatively smalI<br />

pockets. In Quebec, several extensive areas with medium or high populations were<br />

reported between 1974 and 1977. In most <strong>of</strong> Ontario, outbreaks from the previous<br />

decade persisted into the early part <strong>of</strong> the present period, but most populations had<br />

colIapsed by 1972. However, an upsurge in populations was soon noted in southern<br />

Ontario, where severe defoliation <strong>of</strong> larch was reported from 1974 to 1978, particularly<br />

in plantations <strong>of</strong> European larch. In Manitoba, old infestations also persisted into the<br />

early part <strong>of</strong> the period, particularly in the southeastern part <strong>of</strong> the province. A moderate to<br />

severe infestation occurred south <strong>of</strong> The Pas between 1972 and 1975. Only pockets <strong>of</strong><br />

tamarack defoliation were noted in Saskatchewan between 1969 and 1980. Small<br />

pockets <strong>of</strong> defoliation were also reported in Alberta and in the Northwest Territories<br />

during the latter part <strong>of</strong> the period. A few infestations in British Columbia persisted in<br />

the early part <strong>of</strong> the period. The larch sawfly was extremely scarce on western larch until<br />

1976, when a population build-up occurred near Sparwood. This infestation collapsed in<br />

1980.<br />

The origin <strong>of</strong> larch sawfly in North America remains a matter <strong>of</strong> debate. Earlyentomologists<br />

(Fletcher 1885,1906, Fyles 1892, 1906, Hewitt 1912) considered the insect to be a<br />

recent introduction, and a number <strong>of</strong> workers still share this belief (Tumock 1972,<br />

Turnock & Muldrew 1973). The first larch sawfly outbreaks were noted in 1880 on the<br />

369


Pristiphora erichsollii (Hartig). 371<br />

eastern seaboard <strong>of</strong> the United States, and the insect appeared to spread westward<br />

(Coppel & Leius 1955, Nairn et ar. 1962). Reports <strong>of</strong> these early outbreaks indicated<br />

that their effects were particularly devastating. However, some entomologists have<br />

doubted the validity <strong>of</strong> the assumption <strong>of</strong> recent introduction, and feel that the insect has<br />

been here for a long time. Graham (1956) examined growth rings <strong>of</strong>an old tamarack from<br />

northern Michigan and believed that he was able to trace evidence <strong>of</strong> sawfly attack as far<br />

back as 1734. Graham (1930) and Graham & Knight (1965) also reported growth<br />

reductions in tamarack before 1880, which they attributed to larch sawfly defoliation;<br />

however, the periods <strong>of</strong> supposed defoliation do not agree completely. Similarly,<br />

Daviault (1948, 1974) examined some old tamarack from Quebec and noted severe<br />

growth suppression, which he attributed to larch sawfly attack, about 1829 to 1845.<br />

However, Nairn el ar. (I962) pointed out that it is practically impossible to separate the<br />

effects <strong>of</strong> larch sawfly defoliation from growth suppression caused by unfavourable<br />

environmental conditions. Although some <strong>of</strong> their growth records, particularly those for<br />

Isle Pierre, British Columbia, show periods <strong>of</strong> growth suppression prior to the known<br />

occurrence <strong>of</strong> larch sawfly in the area, they did not claim that this suppression was due<br />

to larch sawfly attack. Lejeune & Martin (1948) also examined discs cut from trees up<br />

to 159 years old. They found evidence <strong>of</strong> outbreaks in Manitoba from 1908 to 1915 and<br />

from 1941 to the time <strong>of</strong> examination (1948) but apparently there was no evidence <strong>of</strong><br />

earlier attacks.<br />

Wong (1974) believed that the larch sawfly reached North America during the Miocene<br />

Period via the Bering land bridge. He examined large numbers <strong>of</strong> larch sawfly adults and<br />

identified, on the basis <strong>of</strong> morphological characteristics, five distinct strains. Two<br />

strains occur only in North America, one is confined to Eurasia, and two occur in both<br />

North America and Eurasia, presumably as a result <strong>of</strong> an unintentional introduction<br />

in about 1910.<br />

The early attacks by the larch sawfly (Fyles 1892, Graham 1956) clearly demonstrate<br />

the severe impact that the insect can have on the successful production <strong>of</strong> larch. As a<br />

consequence, long-term population studies were undertaken in southeastern Manitoba<br />

to evaluate the effects <strong>of</strong> various factors on survival <strong>of</strong> the different life stages. These<br />

studies showed that none <strong>of</strong> the "native" agents behaved in a density-dependent manner<br />

and the populations fluctuated unpredictably (Ives 1976). The "native" factors are as<br />

follows. Egg survival is affected by adverse temperatures and several invertebrate<br />

predators, but both types <strong>of</strong> mortality are relatively unimportant. Feeding larvae are<br />

attacked by various vertebrate and invertebrate predators, larvae may be dislodged by<br />

storms, and starvation may cause mortality during severe outbreaks. Some disease<br />

organisms, including fungi, bacteria, microsporidia, and rickettsiae attack feeding larvae,<br />

but their effects are usually negligible. The only "native" parasitoid <strong>of</strong> any consequence in<br />

these studies was the tachinid Bessa harveyi (Tns.), which attacks late-instar larvae.<br />

However, the rates <strong>of</strong> attack were not related to population trends. Long-term sampling<br />

<strong>of</strong> parasitoids <strong>of</strong> larch sawfly in Manitoba and Saskatchewan by the Forest Insect and<br />

Disease Survey also failed to show any relationship between parasitism and population<br />

trends (Ives personal communication). Although parasitism was quite high, it was identical<br />

(44 %) for both increasing and decreasing larch sawfly populations. Elsewhere in <strong>Canada</strong>,<br />

Eclytus ornatus Hlmgr. (which may be a recent introduction from Europe) is a larval<br />

parasitoid <strong>of</strong> some importance in Newfoundland, and the pteromalid Trilneplis klugii<br />

(Ratz.) is occasionally an important cocoon parasitoid in drier sites in British Columbia<br />

(Turnock & Muldrew 1971a). Fully fed larvae dropping to spin cocoons are susceptible to<br />

predation and heat prostration, and some may drown in pools <strong>of</strong> surface water. Survival<br />

in the cocoon phase is affected by a number <strong>of</strong> factors. Those <strong>of</strong> minor importance include<br />

invertebrate predation and fungal diseases. The principal sources <strong>of</strong> mortality in this<br />

phase, apart from introduced parasitoids discussed later, are high water tables and


372 W. G. H. Ives and J. A. Muldrew<br />

predation by small mammals. The combined effects <strong>of</strong> these were key factors in determining<br />

population trends (Ives 1976).<br />

These factors clearly demonstrated the paucity <strong>of</strong> regulating factors in Canadian larch<br />

sawfly populations. Hewitt (191O) had been studying the larch sawfly in England before<br />

his appointment as Dominion Entomologist. He had found the ichneumonid Mesoleills<br />

tenthredinis Morl. to be an effective parasitoid <strong>of</strong> the larch sawfly and, as it was absent<br />

from <strong>Canada</strong>, he made arrangements for its importation and release between 1910 and<br />

1913. The releases in Manitoba were successful and parasitism by M. tenthredinis<br />

gradually increased to nearly 90% by 1927 (Graham 1931) and seemed to be a factor in<br />

terminating the outbreak. However, when the larch sawfly again reached outbreak<br />

proportions, about 1940, it was noted that M. tenthredinis was no longer an effective<br />

parasitoid due to encapsulation <strong>of</strong> the eggs by host blood cells which prevented hatching<br />

(Muldrew 1953). This encapsulation reaction gradually spread across continental North<br />

America, and at present only larch sawfly populations in Newfoundland and cordilleran<br />

British Columbia remain highly susceptible to M. tenthredinis (Bronskill 1960. Carroll<br />

1964). The lack <strong>of</strong> quarantine precautions at the time <strong>of</strong> the initial release was probably<br />

the cause <strong>of</strong> this change in susceptibility. Instead <strong>of</strong> releasing adult parasitoids. the<br />

release in Manitoba was made by placing host cocoons in a bog near Aweme to let the<br />

parasitoids emerge naturally (Hewitt 1917). Because not all the cocoons were parasitized.<br />

sawfly adults were released as well. This seemed harmless enough, as an outbreak was<br />

in progress at the time. However, subsequent research (Maw 1960) indicated that some<br />

<strong>of</strong> these adults probably possessed the encapsulation ability. Their progeny gradually<br />

spread across <strong>Canada</strong>, and have now become the predominant strain in most areas<br />

(Wong 1974).<br />

The decrease in effectiveness <strong>of</strong> M. tenthredinis prompted a renewed interest in<br />

biological control. Redistribution <strong>of</strong> B. harveyi and T. killgii was undertaken (Turnock<br />

& Muldrew 1971a) but the results were disappointing. A decision was therefore made in<br />

1957 to renew the search for suitable candidates in Europe and Japan. A number <strong>of</strong><br />

species were considered, but only two releases were successful; a Bavarian strain <strong>of</strong> M.<br />

tenthredinis and another ichneumonid, Olesicampe benefactor Hinz. In addition, the<br />

masked shrew, Sora cinereus cinereus Kerr, was successfully introduced into Newfoundland<br />

from New Brunswick.<br />

During the collection <strong>of</strong> possible parasitoids in Europe, a number <strong>of</strong> M. tenthredinis<br />

were obtained. Laboratory experiments showed that those from Bavaria were only<br />

slightly encapsulated by their hosts, as were the progeny <strong>of</strong> back-crosses <strong>of</strong> this strain<br />

with M. tenthredinis originating from the earlier introductions. Two releases were made<br />

in Manitoba in 1963 near Hodgson and in 1964 near Rennie (Turnock & Muldrew 1971a).<br />

Increases in parasitism (the two "races" are morphologically indistinguishable) clearly<br />

indicated that the Rennie release was successful. M. tenthredinis from the early release<br />

were more abundant in the Hodgson area and trends following the 1963 release <strong>of</strong> the<br />

Bavarian M. tenthredinis were inconclusive.<br />

A number <strong>of</strong> successful releases <strong>of</strong> O. benefactor were made between 1961 and 1968<br />

(Turnock & Muldrew 1971a). Parasitoids collected in Europe were released in Manitoba<br />

near Pine Falls in 1961 and near Riverton in 1%2 and 1963, and in Saskatchewan near<br />

Crutwell in 1964. Relocations <strong>of</strong> parasitoids collected near Riverton and Pine Falls were<br />

made as follows: Manitoba in 1967 and 1968; Saskat('hewan in 1965; and the maritimes<br />

in 1967 and 1968.<br />

A hyperparasitoid, Mesochorus dimidiatus Hlmgr., has been responsible for limiting<br />

the effectiveness <strong>of</strong> O. benefactor in Europe, and extreme care was exercised in the<br />

initial releases to ensure that none <strong>of</strong> the hyperparasitoids were inadvertently released.<br />

However, M. dimidiatus was recovered from the Pine Falls release area in 1966.<br />

Apparently M. dimidiatus was a rare but widely distributed holarctic species in <strong>Canada</strong><br />

before the release <strong>of</strong> O. benefactor. and has now been recorded in most areas where this


Releases and Recoveries<br />

Olesicampe benefactor<br />

Hinz.<br />

(Hymenoptera:<br />

Ichneumonidae)<br />

Pri.slip/lOra erichsonii (Hartig), 373<br />

parasitoid has been released. Rates <strong>of</strong> hyperparasitism by M. dimidiatus usually increase<br />

rapidly and there is reason to be concerned about the future effectiveness <strong>of</strong> O. benefactor.<br />

The masked shrew, S. cinereus cinereus, was successfully introduced into Newfoundland<br />

in 1958, when 12 females and 10 males were released near St. George's. Most <strong>of</strong> the<br />

individuals in the initial release survived the first winter and their progeny have prospered<br />

and spread and now probably occupy most <strong>of</strong> the suitable habitats throughout the island.<br />

Biological control studies <strong>of</strong> the larch sawfly during the period 1969 to 1980 have<br />

consisted <strong>of</strong> the redistribution <strong>of</strong> O. benefactor and Mesoleius tenthredinis, plus follow-up<br />

studies <strong>of</strong> these parasitoids and <strong>of</strong> the masked shrew in Newfoundland.<br />

Most <strong>of</strong> the redistributions <strong>of</strong> parasitoids involved O. benefactor (Table 95). The<br />

population increase and dispersal <strong>of</strong> O. benefactor are discussed in detail by Muldrew<br />

(in press). He has incorporated additional data that were not analysed by Turnock &<br />

Muldrew (1971a), so that his values for maximum dispersal in 1966 and 1967 are greater.<br />

He reports that the rate <strong>of</strong> dispersal was about 1 km after 4 years, 3 km after 5 years, and<br />

5 km after 6 years. Follow-up sampling until 1974 showed that the rate <strong>of</strong> dispersal<br />

increased dramatically between 1969 and 1971 (Fig. 19). In 1969. the maximum recorded<br />

distance was 87 km (at the Rennie life-table plot). Sampling in 1970 was inadequate for<br />

assessing dispersal but O. benefactor was recovered in northern Minnesota 293 km<br />

south <strong>of</strong> the release point (Kulman et al. 1974, Thompson & Kulman 1976). Extensive<br />

sampling by Muldrew in 1971 detected O. benefactor at Ignace, Ontario, 357 km from<br />

the release point. The maximum detected dispersal distance by 1972 was 476 km.<br />

However, larch sawfly populations began declining throughout the area in 1972, while<br />

hyperparasitism by Mesochorus dimidiatus increased. Apparently these factors influence<br />

the rate <strong>of</strong> spread, as the maximum recorded distance increased by only 64 km, to 540 km,<br />

by 1974.<br />

The dotted lines in Fig. 19 show the approximate maximum rates <strong>of</strong> parasitism at<br />

different distances from the release point. Until 1969 , the points are clustered around the<br />

lines, but the scatter increases for 1971, 1972, and 1974. Part <strong>of</strong> this increased variability<br />

was attributable to the relatively slow spread <strong>of</strong> O. benefactor into southern Manitoba.<br />

Muldrew (in press) surmised that the direction <strong>of</strong> the prevailing winds and abundance <strong>of</strong><br />

tamarack may have been the main reasons for the slow dispersal into this area. However<br />

by 1974 the rates <strong>of</strong> parasitism in southern Manitoba had increased considerably, and<br />

there was already some indication <strong>of</strong> a relative decline in rates <strong>of</strong> parasitism elsewhere,<br />

possibly due to hyperparasitism by M. dimidiatus. Although not shown in Fig. 19 the<br />

rates <strong>of</strong> M. dimidiatus parasitism increased dramatically from 1969 to 1974. In 1969,<br />

about 80% <strong>of</strong> O. benefactor near the release point were parasitized by M. dimidiatus. In<br />

1971, high rates <strong>of</strong> hyperparasitization had extended to nearly 200 km from the release<br />

point. By 1972, the high levels <strong>of</strong> M. dimidiatus attack had extended to over 300 km.<br />

Muldrew (personal communication) indicatedM. dimidiatus probably transferred to O.<br />

benefactor at a number <strong>of</strong> locations. The apparently rapid rate <strong>of</strong> dispersal between 1969<br />

and 1971, and the variable levels <strong>of</strong> hyperparasitism from 1971 to 1974 both agree with<br />

this hypothesis. There was little additional increase <strong>of</strong> M. dimidiatus parasitism in 1974,<br />

but the apparent decrease in O. benefactor parasitism is probably attributable to the<br />

hyperparasitoid.<br />

Information on increases in parasitism and dispersal from other release points is<br />

rather limited. A release <strong>of</strong> 694 males and 737 females was made near The Pas, Manitoba


374 W. G. H. (ves and J. A. Muldrew<br />

Table 95<br />

Species and province<br />

Redistribution releases and recoveries <strong>of</strong> Olesicampe benefactor Hinz and Mesoleius<br />

tenthredinis Morl. against Pristiphora erichsonii (Htg.)<br />

Location <strong>of</strong> release Numbers released<br />

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

Year Origin Lat. long. Males Females Total recovery<br />

"N OW<br />

Olt!sicampt! btlU!factor Hinz<br />

Prince Edward Island 1971 Nova Scotia 46001' 62°44' 157 117 274 1974<br />

1972 Nova Scotia 46"29' 63°58' 198 189 387<br />

Nova Scotia 1973 Nova Scotia 45°14' 62"49' 393 457 8SO<br />

1975 Nova Scotia 45°43' 63"09' 63 1976<br />

1975 Nova Scotia 45°16' 63°18' 20 1976<br />

1976 Nova Scotia 45"21' 61°54' 45 1977<br />

1976 Nova Scotia 45°19' 62°15' 69 1977<br />

Ontario 1976 Minnesota 79"18' 44004' 182 213 395 1977<br />

1978 Manitoba 75°11' 45°12' 30 32 62<br />

Manitoba 1969 Manitoba 49"48' 97"08' 60 53 113 1974<br />

1970 Man. and Sask. 49"19' 96000' 421 443 864 1970<br />

1971 Manitoba 50003' 96°13' 245 262 S07 Already present<br />

1972 Manitoba 49"19' 96000' 75 74 149<br />

Alberta 1972 Manitoba 54°43' 110004' 1283"<br />

1972 Manitoba 54"27' 113°58' 1139" 1973<br />

1972 Manitoba 55001' 119"06' 469- 1973<br />

1973 Alberta 54·30' 112°57' 118 122 240 1974<br />

1975 Manitoba 53"32' 117'04' 51 86 137 1975<br />

1981 Alberta 59"55' 111°43' 36<br />

Northwest Territories 1m Manitoba 60")6' 116006' 856-<br />

1981 Alberta 60"01' 112"07' 332 1981<br />

1981 Alberta 60"01' 110058' 30<br />

British Columbia 1980 Alberta 49"42' 114·55' 170<br />

1980 Alberta 49"29' 115004' 145<br />

1980 Alberta 49"38' 114·56' 150<br />

Mtsoltius Itnlhrt!dinis Morley<br />

Manitoba 1970 Man. and Sask. 4919' 96000' 63 85 148<br />

1971 Manitoba 49"19' 96000' 98 170 268<br />

1m Manitoba 49"19' 96000' 9 42 51<br />

• Estimates only - "small" cocoons were placed in the field, adult parasitoid emergence holes were counted<br />

and proportions <strong>of</strong> O. benefactor and Mesochorus'dimidiatus were based on reared samples.<br />

Meso/eius tenthred;n;s<br />

Morley<br />

(Hymenoptera:<br />

Ichneumonldae)<br />

in 1968. O. benefactor had dispersed 18 km by 1973 and 57 km by 1975,7 years after the<br />

release. Another release, consisting <strong>of</strong> 51 males and 86 females, was made near Obed<br />

Lake, Alberta, in 1975. O. benefactor parasitism near the release area reached 88% by<br />

1979 and 93 % by 1980. Seventeen per cent <strong>of</strong> the larch sawfly larvae in a bog 15 km east <strong>of</strong><br />

the release site were parasitized by O. benefactor in 1980. None had been found in this bog<br />

in 1979.<br />

Dispersal from The Pas and Obed Lake releases, 18 km and 15 km in 5 years, is<br />

considerably greater than the 3 km in 5 years recorded for the initial release at Pine Falls.<br />

This may simply be due to a different set <strong>of</strong> environmental conditions, but it could<br />

indicate that O. benefactor has become better adapted to the Canadian environment.<br />

Redistributions are shown in Table 95.


376 W. G. H. Ives and J. A. Muldrew<br />

Sorex cinereus<br />

cinereus Kerr<br />

(lnsedivora: Sorlddae)<br />

Evaluation <strong>of</strong> Control Attempts<br />

The masked shrew, introduced into Newfoundland in 1958, has continued to spread.<br />

Survey reports for 1969 and 1970 (Anon. 1970-79) mentioned the concern <strong>of</strong> local<br />

residents as high initial numbers were noted in each area: by 1971 the shrew had spread<br />

over 95% <strong>of</strong> the island. Populations on the island, excluding the first 2 or 3 years' data,<br />

averaged about eight shrews per hectare, which seems to be unusually high. However,<br />

the 18-year average populations <strong>of</strong> the masked shrew in the Rennie life-table plot (Ives<br />

1981) was 5.2/ha. If one includes populations <strong>of</strong> the closely related arctic shrew, Sora<br />

arcticus Kerr, which averaged 1.8/ha, the total <strong>of</strong> 7.01ba is comparable. Populations <strong>of</strong><br />

the masked shrew in Newfoundland therefore seem to have stabilized, and can be<br />

expected to fluctuate in the future in response to environmental conditions.<br />

O. benefactor and larch sawfly populations have probably not yet reached an equilibrium,<br />

even at the original release point near Pine Falls. As shown in Fig. 19, the level <strong>of</strong> O.<br />

benefactor parasitism at or near the release point reached 90% in 1967. It remained at or<br />

near this level until 1972, but larch.sawfly populations collapsed in 1973, presumably as a<br />

result <strong>of</strong> the continued high rates <strong>of</strong> parasitism (rves 1976). No larvae or cocoons were<br />

obtained near the plot in 1973 or 1974, although intensive sampling was conducted<br />

(Muldrew in press). Detailed studies were discontinued after 1974, and further trends<br />

are based on limited sampling. In 1977, a collection <strong>of</strong> 72 fourth- and fifth-instar larvae<br />

was made near the original release site at Pine Falls. None was parasitized by O.<br />

benefactor. In 1978, collections <strong>of</strong> larvae from the same area yielded 1730 cocoons.<br />

Parasitism by O. benefactor was 1.5%, hyperparasitism by Mesochorus dimidiatus was<br />

86%. A total <strong>of</strong> 237 third- to fifth-instar larvae collected 8 km north <strong>of</strong> the release point in<br />

1978 contained no O. benefactor, but 19% <strong>of</strong> 424 fourth- and fifth-instar larvae from an<br />

area 8 km south <strong>of</strong> the release point were parasitized. NinetY7three per cent <strong>of</strong> these O.<br />

benefactor larvae were parasitized by M. dimidiatus. In 1980, a total <strong>of</strong> 535 cocoons<br />

was obtained from the same location; 15% <strong>of</strong> these contained O. benefactor, <strong>of</strong> which<br />

75% were parasitized by M. dimidiatus. No O. benefactor were present in 150 cocoons<br />

obtained by rearing late-instar larvae collected in 1980 near the Seddon's Comer lifetable<br />

plot, about 70 km south <strong>of</strong> the Pine Falls release point.<br />

Although the above data are rather fragmented, there is some evidence that O.<br />

benefactor is in danger <strong>of</strong> local extinction over a fairly wide area. Should this happen, it<br />

is possible that the larch sawfly may increase in abundance if environmental conditions<br />

are favourable. No population sampling <strong>of</strong> the larch sawfly was done in the Pine Falls<br />

release area in 1980, but observations by Ives indicated that this population increase<br />

may already be taking place. It remains to be seen whether or not O. benefactor will be<br />

able to prevent an outbreak in spite <strong>of</strong> the adverse influence <strong>of</strong> M. dimidiatus. O.<br />

benefactor is host specific, but M. dimidiatus is not (Pschorn-Walcher & Zinnert 1971).<br />

Because <strong>of</strong> this, one would expect M. dimidiatus to be able to persist on alternate hosts,<br />

thus obtaining a "head start" on any re-invading O. benefactor. Under these circumstances,<br />

M. dimidiatus would probably be able to control O. benefactor, thus permitting larch<br />

sawfly populations to reach outbreak proportions. However, a crude model developed<br />

by Ives (1976) indicated that the two parasitoids should reach an equilibrium with each<br />

other and with the larch sawfly. According to the model, the larch sawfly populations<br />

would fluctuate in response to density-independent environmental factors, but should not<br />

reach outbreak proportions in the presence <strong>of</strong> O. benefactor, even when M. dimidiatus is<br />

present. It is too early to tell which hypothesis is correct.<br />

The impact <strong>of</strong> the masked shrew on larch sawfly populations in Newfoundland has not<br />

been assessed in detail (Warren 1971). The fact that larch sawfly infestations reached


i'risliphora erichsonii (Hartig). 377<br />

outbreak proportions late in the decade 1971-1980 clearly indicates that the shrews<br />

were not able to control the insect. However, as Warren (1971) indicated, "the shrew has<br />

been a valuable predator <strong>of</strong> larch sawfly, and, despite some evidence <strong>of</strong> reduced<br />

parasitism, there has been an important net gain in cocoons destroyed". Ives (1976)<br />

showed that mortality during the cocoon and adult stages was largely responsible for<br />

determining population trends. Small mammal predation is <strong>of</strong>ten a major component <strong>of</strong><br />

this mortality, thus agreeing with Warren's assessment <strong>of</strong> the effect <strong>of</strong> the masked shrew<br />

in Newfoundland.<br />

Although there is some indication that larch sawfly popUlations in and around the Pine<br />

Falls release site may now be on the increase, there has been a period <strong>of</strong> several years<br />

when this insect was extremely difficult to find. It was probably not as rare as in Europe,<br />

but it was much less common than it has been in recent memory. Two recent papers<br />

illustrate just how rare the larch sawfly can be in alpine areas <strong>of</strong> Europe. Lovis (1975)<br />

collected only six larch sawfly (out <strong>of</strong> a total <strong>of</strong> 2816 larch insects) during a 3-year study<br />

<strong>of</strong> insects attacking native European larch in Switzerland. Auer (1971) studied the<br />

population dynamics <strong>of</strong> the larch bud moth, Zeiraphera diniana Gn., in the Upper<br />

Engadine Valley in Switzerland between 1949 and 1968. He presented population curves<br />

for eight larch insects for the period 1952 to 1968. The larch sawfly was apparently too<br />

rare to be included.<br />

The infestation patterns <strong>of</strong> the larch sawfly have been discussed by Turnock (1972).<br />

He did not consider the effects <strong>of</strong> O. benefactor, but with slight changes his discussion is<br />

still valid. He recognized three types <strong>of</strong> life systems: I) the stable latent type with a<br />

diverse environment and a rich parasitoid complex, characteristic <strong>of</strong> infestations on larch<br />

in alpine Europe; 2) the stable permanent type with a paucity <strong>of</strong> specific parasitoids,<br />

characteristic <strong>of</strong> many North American outbreaks before 1920; and 3) the temporary type<br />

with an intermediate parasitoid complex which includes one (or two) very effective larval<br />

parasitoid(s). characteristic <strong>of</strong> European plantings. infestations on western larch in<br />

British Columbia and on tamarack in the rest <strong>of</strong> <strong>Canada</strong> following the first successful<br />

introduction <strong>of</strong> an ichneumonid parasitoid between 1910 and 1913 (and in southeastern<br />

Manitoba following the successful introduction <strong>of</strong> O. benefaclor there in 1961). As this<br />

discussion <strong>of</strong> life systems illustrates, entomologists have been able to achieve at least<br />

partial biological control <strong>of</strong>the larch sawfly in North America. To continue this control, it<br />

may be necessary to seek ways in which biotic agents in the environment can be manipulated.<br />

Turnock & Muldrew (1971b) and Turnock et al. (1976) recognized four ways in<br />

which insect parasites or predators might be manipulated in order to help achieve<br />

biological control: 1) colonization releases; 2) inoculative releases; 3) inundative releases;<br />

and 4) enhancement <strong>of</strong> biotic agents by environmental manipulation. In the case <strong>of</strong> the<br />

larch sawfly, the last three approaches <strong>of</strong>fer the most promise. Because <strong>of</strong> the presence<br />

<strong>of</strong> M. dimidiatus, any releases <strong>of</strong> O. benefactor would need to be relatively large if they<br />

were to be effective. Releases intermediate in size between inoculative and inundative.<br />

say 2 000-3000 mated females, should be able to control infestations in plantations in 2<br />

or 3 years, if released when defoliation first becomes noticeable. However, this approach<br />

would necessitate a permanent facility capable <strong>of</strong> producing parasitoids on demand.<br />

Reliance on field populations would be too uncertain, especially as M. dimidiatus will<br />

almost certainly reduce the numbers <strong>of</strong> O. benefactor adults obtainable. The Bavarian<br />

strain <strong>of</strong> Mesoleius tenthredinis could probably be managed in a similar manner. Both<br />

species might even be able to sustain themselves in the field, and thus afford more or less<br />

permanent protection. In Manitoba O. benefactor appeared to suppress M. tenthredinis<br />

by sheer numbers. but this seems unlikely to happen again.<br />

Enhancement <strong>of</strong> environmental conditions might be a particularly effective method<br />

for increasing the effectiveness <strong>of</strong> avian and small mammalian predators in plantations.<br />

Buckner (1971) discussed this situation as it related to forest insects in general. Birds<br />

might be encouraged to nest in plantations by providing suitable nesting sites. Bruns


378 W. G. HIves and J. A. Muldrew<br />

Recommendations<br />

(1960) reported that this approach appeared to be effective in controlling pine insects in<br />

Germany. In the case <strong>of</strong> small mammals, research would have to be conducted to ensure<br />

that voles do not attack larches. Neither <strong>of</strong> us has observed evidence <strong>of</strong> this, but some<br />

tree species are severely damaged by voles. If high vole populations are not a threat to<br />

the trees it should be possible to enhance the environment by providing adequate cover,<br />

either in the form <strong>of</strong> slash or between-row vegetation, as several workers have shown<br />

that mouse and shrew populations are highest in areas with adequate cover (Morris<br />

1955, Beer 1961, Coulianos & Johnels 1962). Any such manipulation would have to be<br />

done carefully, to avoid creating an environment favouring snowshoe hares. However,<br />

it should be possible to improve the environment <strong>of</strong> the plantations as habitats for voles<br />

and shrews but not hares. This would increase the amount <strong>of</strong>small mammal predation,<br />

which should reduce the frequency <strong>of</strong> larch sawfly outbreaks.<br />

In low-lying situations it may be possible to flood the plantations at critical times in the<br />

larch sawfly's life cycle, as suggested by Buckner (1971). Hooding the plantations for<br />

short periods in the fall, soon after cocoon formation, or in the spring when the insects<br />

are in the pupal stage would have a marked effect on larch sawfly survival (Lejeune et 01.<br />

1955) but would not seriously damage the tamarack trees if carefully timed. It would also<br />

adversely affect small mammal populations, but this could be minimized by careful<br />

timing. If flooding could be done just before parasitoids are released, it might<br />

increase their effectiveness considerably, as the initial rate <strong>of</strong> parasitism would be<br />

higher.<br />

Finally, should serious infestations develop that may endanger the trees in plantations,<br />

it may still be possible to control the larch sawfly without unduly affecting the other<br />

biotic agents. In the case <strong>of</strong> M. tenthredinis, spraying with a larval insecticide (preferably<br />

one that does not kill adult Hymenoptera) when most <strong>of</strong>the larch sawfly larvae are in the<br />

first to third instars would reduce the sawfly populations but should not have an undue<br />

adverse effect on the parasitoids, as they attack mainly fourth- and fifth-instar larvae.<br />

Because <strong>of</strong> the prolonged period <strong>of</strong> larch sawfly emergence it would not be feasible to kill<br />

all the larvae, but it should be possible to reduce populations appreciably and thus<br />

enhance the effectiveness <strong>of</strong> M. tenthredinis. The use <strong>of</strong> chemicals in the presence <strong>of</strong> O.<br />

benefactor would be more difficult. Probably the most productive approach would be to<br />

reduce sawfly populations only to the extent that defoliation would not exceed 70%. This<br />

would save the trees, but should not have an unduly adverse effect on the parasitoids.<br />

The Bavarian strain <strong>of</strong> M. tenthredinis was released near Rennie, Hodgson, and St.<br />

Labre. It is known to have become established at Rennie, but how far it spread before<br />

being overwhelmed by O. benefactor is not known. The extremely low density <strong>of</strong> larch<br />

sawfly in the vicinity <strong>of</strong> the Rennie plot for several years after its establishment may<br />

mean that its numbers were markedly reduced, if not eliminated. No follow-up studies<br />

were conducted in the St. Labre area, so it is not known if the Bavarian strain (from the<br />

Rennie plot) became established or not.<br />

It is therefore recommended that the Bavarian strain <strong>of</strong> M. tenthredinis be released in<br />

the Pine Falls area as soon as larch sawfly populations in the Bavarian plantations are<br />

high enough to make collection feasible. The results <strong>of</strong> the release at the Rennie plot<br />

looked extremely encouraging and a second attempt to establish this strain seems<br />

warranted. In Europe, M. tenthredinis is the most important insect parasite <strong>of</strong> the larch<br />

sawfly (Pschom-Walcher & Zinnert 1971) and there seems to be a reasonable chance <strong>of</strong><br />

the Bavarian strain assuming this position in <strong>Canada</strong>, now that Mesochorus dimidiatus<br />

seems destined to limit the effectiveness <strong>of</strong> Q. benefactor.<br />

O. benefactor has been widely redistributed in various parts <strong>of</strong> <strong>Canada</strong>, from the<br />

maritimes to British Columbia. There does not seem to be much point in continuing with


380 W. G. H. Ives and J. A. Muldrew<br />

Maw, M.G. (1960) Notes on the larch sawfly, Pristiphora erichsonii (Htg.) (Hymenoptera: Tenthredinidae), in Great Britain. Entomologist's<br />

Gazette 11,43-49.<br />

McGugan, B.M.; Coppel. H.C. (1962) Biological control <strong>of</strong> forest insects - 1910-1958. In: A review <strong>of</strong> the biological control allempts<br />

against insects and weeds in <strong>Canada</strong>. Commonwealth Institute <strong>of</strong> Biological Control Technical Communication<br />

2,35-127.<br />

Morris, R.F, (1955) Population studies on some small forest mammals in eastern <strong>Canada</strong>. Journal <strong>of</strong> Mammalogy 36,21-35.<br />

Muldrew, J.A. (1953) The natural immunity <strong>of</strong> the larch sawRy (Pristiphora erichsonii (Htg.» to the introduced parasite Mesoleius<br />

tenthredinis Morley, in Manitoba and Saskatchewan. Canadian Journal <strong>of</strong> Zoology 31,313-332.<br />

Muldrew, J.A. (In press) Dispersal <strong>of</strong> Olesicampe benefactor Hinz, an introduced parasite <strong>of</strong> the larch sawfly. Canadian Forestry Service.<br />

Northern Forest Research Centre. Edmonton. Infonnation Report.<br />

Nairn. L.D.; Reeks. W.A.; Webb. F.E.; Hildahl, V. (1962) History <strong>of</strong> larch sawfly outbreaks and their effects on tamarack stands in<br />

Manitoba and Saskatchewan. Canadwn Entomologist 94.242-255.<br />

Pschom-Walcher, H.; Zinnen. K.D. (1971) Investigations on the ecology and natural control <strong>of</strong> the larch sawRy (Pristiphora erichsonii Htg.,<br />

Hym.: Tenthredinidae) in central Europe. Pan II: Natural enemies: their biology and ecology, and their<br />

role as monality factors in P. erichsonii. Commonwealth Institute <strong>of</strong> Biological Control Technical<br />

Bulletin 14,1-50.<br />

Thompson, L.C.; Kulman, H.M. (1976) Parasite complex <strong>of</strong> the larch sawRy in Minnesota. Environmental Entomology 5.1121-1127.<br />

Tumock, W.J. (1972) Geographical and historical variability in population pallems and life systems <strong>of</strong> the larch sawRy (Hymenoptera:<br />

Tenthredinidae). Canadian Entomologist 104,1883-1900.<br />

Turnock, W.J.; Muldrew, J.A. (1971a) Pristiphora erichsonii (Hanig), larch sawfly (Hymenoptera: Tenthredinidae). In: Biological control<br />

programmes against insects and weeds in <strong>Canada</strong>, 1959-1968. Commonweahh Institute <strong>of</strong> Biological<br />

Control Technical Communication 4,175-194.<br />

Tumock, W.J.; Muldrew, J.A. (1971b) Parasites. In: Toward integrated control. US Depanment <strong>of</strong> Agriculture Forest Service Research<br />

Paper NE-I94. Nonheastem Forest Experiment Station, Upper Darby, Pennsylvania. pp. 59-87.<br />

Tumock, W.J.; Muldrew, J. A. (1973) Characteristics <strong>of</strong> Bessa harveyi (Diptera: Tachinidae) suggesting the historic introduction <strong>of</strong> the larch<br />

sawfly to Nonh America. Manitoba Entomologist 6,49-53.<br />

Tumock, W.J.; Taylor, K.L.; Schroder, D.; Dahlsten, D.L. (1976) Biological control <strong>of</strong> pests <strong>of</strong> coniferous forests. In: Huffaker, C.B.;<br />

Messenger, P.S. (Eels.). Theory and practice <strong>of</strong> biological control. New York; Academic Press, pp.<br />

289-311.<br />

Warren, G.L. (19ul) Introduction <strong>of</strong> the masked shrew to imrrove control <strong>of</strong> forest insects in Newfoundland. Proceedings <strong>of</strong> the Tall Timbers<br />

Conference on Ecologica Animal Control by Habitat Management 2,185-202.<br />

Wong, H.R. (1974) The identification and origin <strong>of</strong> the strains <strong>of</strong> the larch sawRY' Pristiphora erichsonii (Hymenoptera: Tenthredinidae), in<br />

North America. Canadian Entomologist HI6.1121-1131.


Pest Status<br />

Background<br />

Chapter 65<br />

Pristiphora geniculata (Htg. ), Mountain­<br />

Ash Sawfly (Hymenoptera: Tenthredinidae)<br />

F.W. QUEDNAU<br />

The mountain-ash sawfly, Pristiphora geniculata (Hartig), is a major defoliator <strong>of</strong><br />

mountain-ash, mainly Sorbus americana Marsh, and S. aucuparia L. These tree species<br />

have little industrial importance but are planted widely as shade and ornamental trees.<br />

The mountain-ash sawfly was accidentally introduced into the United States in 1926<br />

(Schaffner 1936) and was first recorded in <strong>Canada</strong> in southern Quebec in 1934 (Petch<br />

1935). Its present Canadian range is from Newfoundland to southwestern Ontario (Fig.<br />

20). There has never been a serious outbreak. The sawfly causes noticeable and occasionaJly<br />

complete defoliation <strong>of</strong> mountain-ash, but this seldom causes mortality <strong>of</strong> the<br />

tree. To owners <strong>of</strong> such trees and to municipal park authorities, the insect is a considerable<br />

nuisance and <strong>of</strong>ten requires localized control by chemical methods. An account <strong>of</strong> the<br />

biology <strong>of</strong> mountain-ash sawfly in eastern <strong>Canada</strong> was given by Forbes & Daviault<br />

(1964). P. geniculata usually has one generation a year, but there may be a partial second<br />

generation depending on the climate.<br />

The decision to embark on a biological control programme in Quebec against P. geniculata<br />

was determined by the policy to reduce chemical control in urban forestry and by the<br />

success <strong>of</strong> biological control in Manitoba against the closely related larch sawfly, P.<br />

erichsonii (Htg.), on tamarack, Larix laricina (Du Roi) K. Koch (Muldrew 1967, 1973).<br />

Data furnished by the CIBC European Station suggested that the ichneumonid wasp<br />

Olesicampe geniculatae Quednau & Lim, would be a primary candidate for introduction<br />

into <strong>Canada</strong>. Another parasitoid species under consideration was the ichneumonid wasp<br />

Rhorus sp. No.3.<br />

During the first years <strong>of</strong> attempts to colonize these parasitoid species in Quebec,<br />

infestation levels <strong>of</strong> mountain-ash sawfly were very low. Furthermore, in 1972 O. geniculatae<br />

was not found in sufficient numbers in Europe to warrant a shipment to <strong>Canada</strong>.<br />

However, in 1973, heavy infestations <strong>of</strong> P. geniculata occurred in several localities near<br />

Quebec City. Because <strong>of</strong> the scarcity in Europe <strong>of</strong> P. geniculata, laboratory massrearings<br />

<strong>of</strong> the sawfly were carried out in Quebec in order to obtain host material that<br />

could be sent to Europe to increase the host populations there. The rearing techniques<br />

were essentially the same as those described for the larch sawfly by Heron & Drouin<br />

(1969). In 1974 the Laurentian Forest Research Centre sent 6 000 cocoons <strong>of</strong> Canadian P.<br />

geniculata to CIBC in Switzerland and another 30 000 cocoons in 1976. This insect<br />

material was reared to the adult stage in Europe and the sawfly females were placed in<br />

plastic bags tied over branches <strong>of</strong> mountain-ash trees in the Waldviertel area in Austria,<br />

where O. geniculatae and other parasitoid species <strong>of</strong> the mountain-ash sawfly were known'<br />

to occur in significant numbers during the previous years. As a result, in 1977 large<br />

numbers <strong>of</strong> O. geniculatae and Rhorus sp. No.3 were received in Quebec.<br />

381


Releases and Recoveries<br />

Rhorus sp. No.3<br />

(Hymenoptera:<br />

[chneumonidae)<br />

Table 96<br />

0IeSaHnpe genJculBtae<br />

Quednau & Lim<br />

(Hymenoptera:<br />

[chneumonidae)<br />

The utilization <strong>of</strong> the parasitoid material is summarized in Table 96.<br />

Pristiphora geniclllata (Htg.), 383<br />

Parasitoids utilized for releases against mountain-ash sawfly, Pristiphora geniculoJa (Htg.),<br />

at Beaumont, Quebec (Iat. 46°50'N, long. 71002'W)<br />

Species Year Origin<br />

Rhorus sp. No.3 1973 Austria<br />

1974 Switzerland<br />

and Austria<br />

1975 Austria<br />

1976 Austria<br />

1977 Austria<br />

Olesicampe geniclliatae 1974 Switzerland<br />

and Austria<br />

1975 Austria<br />

1976 Austria<br />

1977 Austria<br />

No. and sex <strong>of</strong><br />

parasitoids received<br />

Male Female<br />

5<br />

29<br />

12<br />

67<br />

35<br />

22<br />

5<br />

32<br />

16<br />

83<br />

35<br />

16<br />

2 3<br />

92 117<br />

315 551<br />

This is a solitary endoparasitoid, probably specific to P. geniculata. The egg is laid<br />

through the ocellus into the head capsule <strong>of</strong> the host larva. This wasp is closely related to<br />

R. iapponicus Roman, a parasitoid <strong>of</strong> the larch sawfly, the biology <strong>of</strong> which was<br />

described by Pschorn-Walcher & Zinnert (1971). Rhorus sp. No.3 was received in<br />

Quebec from 1973 to 1977. Cage releases were made at Beaumont nurseries. A male was<br />

recovered in the field in 1978, but the parasitoid apparently did not become permanently<br />

established. It is difficult to mate this insect in captivity.<br />

This is also an internal solitary parasitoid that prefers the first- and second-instar larvae<br />

<strong>of</strong> the mountain-ash sawfly for attack. It is specific to this host. In the early colonization<br />

work in Quebec only small numbers <strong>of</strong> the parasitoid were available and shipments from<br />

Europe were <strong>of</strong>ten not well synchronized with host populations in Quebec. No recoveries<br />

were obtained from these first releases. Major implantations <strong>of</strong> this parasitoid were<br />

made in 1977 at Beaumont nurseries. The female parasitoids were mated in the laboratory<br />

and released in batches in large tents <strong>of</strong> woven plastic screening built over mountain-ash<br />

trees, in which host densities were artificially increased by adding young colonies <strong>of</strong> P.<br />

genicuiata larvae collected from other areas. Small bottles <strong>of</strong> water were fastened to the<br />

branches <strong>of</strong> the trees in the tents. The host material was exchanged and new parasitoids<br />

added each day. The parasitized P. geniclliata larvae were brought to the laboratory,<br />

given plenty <strong>of</strong> mountain-ash foliage that was kept fresh in water bottles, and reared to<br />

about fourth instar. This method excluded natural predators. The sawfly larvae were<br />

then returned to the field and placed in wire baskets (100 per basket) containing some<br />

foliage. The baskets were tied to the branches <strong>of</strong> mountain-ash trees in the area chosen<br />

for implantation. Parasitized larvae eventually fell to the ground and formed cocoons.<br />

Successful establishment <strong>of</strong> this parasitoid depends on massive initial implantations<br />

to give the parasitoid a good chance to multiply in the field. A female O. genicu/atae can<br />

lay up to 400 eggs and live up to 3 weeks in the laboratory at 22°C and 75% relative


384 F. W. Quednau<br />

Evaluation <strong>of</strong> Control Attempts<br />

humidity (F.W. Quednau unpublished). The biology <strong>of</strong> this species is essentially the<br />

same as that reported for o. benefactor Hinz by Pschorn-Walcher & Zinnert (1971).<br />

O. geniculatae may have a partial second generation each year in the field.<br />

Whereas the colonization <strong>of</strong> Rhorus sp. No.3 was unsuccessful, O. geniculatae became<br />

firmly established at Beaumont nurseries, and within 4 years parasitized a large proportion <strong>of</strong><br />

hosts. A brief history <strong>of</strong> this event follows.<br />

At the beginning <strong>of</strong> massive implantations in 1m, population densities <strong>of</strong> P. geniculata<br />

were medium to heavy, and 1 year later were light to medium. In the fall <strong>of</strong> 1978 it was<br />

found from dissections <strong>of</strong> sawfly cocoons that the percentage <strong>of</strong> parasitism at the release<br />

centre had dropped to 25-30%, compared to 60-75% in 1977, as a result <strong>of</strong>the dilution<br />

<strong>of</strong> the parasitoid population in the area. It was estimated that within 1 year O. geniculatae<br />

spread about 500 m from the original release centre. In 1979 infestations by the mountainash<br />

sawfly were only light at Beaumont. Of30soil samples <strong>of</strong> 0.1 m l each, taken near the<br />

original release point <strong>of</strong> the parasitoid, only 10 contained living cocoons (range 1-5) and<br />

parasitism was about 73%, much higher than in 1978. Other sampling plots about 500 m<br />

from the release point yielded 30-35% parasitism in 1979. In 1980, in the same locality,<br />

dissections <strong>of</strong> mountain-ash sawfly larvae that were still in colonies (about third instar)<br />

were made. The infestation level was from 700 to 1 250 young colonies <strong>of</strong> the sawfly per<br />

hectare. Percentage parasitism obtained from these dissections was 88.7-93.5% in all the<br />

plots containing mountain-ash in the nursery (up to 500 m radius from the release centre).<br />

No second generation <strong>of</strong> P. geniculata was observed in 1980. From an evaluation <strong>of</strong> host<br />

cocoons collected in the fall <strong>of</strong> 1980 it was found that the number <strong>of</strong> P. geniculata cocoons<br />

had further diminished as compared with 1979 and few were unparasitized.<br />

Preliminary indications for 1981 are that O. geniculatae has spread about 30 km from<br />

the original release site and could be recovered at Ste-Foy, Charlesbourg, La Durantaye,<br />

St-Charles de Bellechasse, St-Michel de Bellechasse, St-Etienne de Lauzon, and Chutes<br />

Mt. Ste-Anne where parasitism was 81, 24, SO, 30, 82,70, and 4% respectively. A quantitative<br />

evaluation <strong>of</strong> the impact <strong>of</strong> the parasitoid on its host population has not been carried out<br />

so far. However, preliminary measures <strong>of</strong> pest numbers in consecutive years in the<br />

original release site at Beaumont indicate a drastic decline <strong>of</strong> the host populaton (from<br />

500 colonieslha in 1980 to an estimated 30 colonieslha in 1981), which must be attributed<br />

primarily to the action <strong>of</strong> the parasitoid. Comparative population counts <strong>of</strong> P. geniculata<br />

at St-Edouard de Frampton, where the environmental structure is very similar to that at<br />

Beaumont, but is situated outside the action radius <strong>of</strong> o. geniculatae, indicate very high<br />

levels <strong>of</strong> infestation in 1981.<br />

The cost <strong>of</strong> chemical insecticide treatment against P. geniculala would be at least<br />

$2OOlha. For this reason most owners <strong>of</strong> property with mountain-ash do not use chemical<br />

treatments. The mass production <strong>of</strong> O. geniculatae would need extensive manual labour<br />

and costs cannot be calculated accurately at present. However, as the parasitoid spreads<br />

on its own, its implantation could be highly beneficial.<br />

During the evaluation work in summer 1980, a large number <strong>of</strong> the larvae <strong>of</strong><br />

O. geniculatae found in the hosts were dead or appeared sick. At this stage it was decided<br />

to rear adults from P. geniculata cocoons collected from Beaumont in order to determine<br />

the percentage parasitism by O. geniculatae and to find out whether a hyperparasitoid was<br />

present. Emergence <strong>of</strong> the insects in 1981 revealed up to 90% total parasitism, but about<br />

60% <strong>of</strong> the parasitoids obtained were Mesochorus globu/alorThnb. This latter species is a<br />

hyperparasitoid with a wide distribution in <strong>Canada</strong>. It attacks various other species <strong>of</strong><br />

Olesicampe associated with nematine sawflies.


Blank Page<br />

386


Pest Status<br />

Chapter 66<br />

Rhyacionia buoliana (Schiff.),<br />

European Pine Shoot Moth<br />

(Lepidoptera: Tortricidae)<br />

P.D. SYME<br />

The European pine shoot moth, Rhyacionia buoliana (Schiff.), has been known as a<br />

destructive pest <strong>of</strong> hard pine, Pinus spp., plantations in North America since it was<br />

identified in 1914 (Busk 1914). Its history and biology have been the subjects <strong>of</strong> extensive<br />

investigations prior to 1969 and are well summarized by McGugan & Coppel (1962),<br />

Pointing & Green (1962), Miller (1967), and Syme (1971). The status <strong>of</strong> R. buoliana has<br />

been under continual surveillance and has been recorded in the Annual Reports <strong>of</strong> the<br />

Forest Insect and Disease Survey. The known distribution is shown in Fig. 21 for eastern<br />

<strong>Canada</strong> and British Columbia.<br />

The European pine shoot moth has one generation per year and adults appear in the<br />

field over a 4·week period that begins between the first and third weeks in June. Eggs are<br />

laid singly or in small groups on the twigs or needle sheaths <strong>of</strong> the new growth. Incubation<br />

requires 2 weeks. First- and second-instar larvae mine needle bases on the new growth.<br />

Later they migrate to the new buds where they feed before overwintering as half-grown<br />

larvae within the bud. Feeding is resumed in the spring and is extended to additional<br />

buds before pupation takes place in May.<br />

R. buoliana is primarily a pest <strong>of</strong> pine plantations. In North America, red pine, Pinus<br />

resinosa Ait., is the most seriously damaged. Scots, Austrian, ponderosa, and ornamental<br />

Mugho pines, P. sy/vestris L., P. nigra Arnold, P. ponderosa Laws., and P. mugo<br />

Turra, respectively, are moderately susceptible; whereas pitch, Virginia, jack, and eastern<br />

white pines, P. rigida Mill., P. virginiana, P. banksiana Lamb., and P. strobus L., respectively,<br />

are relatively resistant. The needle-mining larvae cause a browning <strong>of</strong> the foliage,<br />

but this apparently has little effect on the vigour <strong>of</strong> the host. Feeding on the terminal bud<br />

causes serious injury to the main stem form and repeated destruction <strong>of</strong> the entire<br />

terminal cluster during the summer months initially causes pr<strong>of</strong>usion <strong>of</strong> shoot growth<br />

(witch's broom) and eventually a spike top when the leader dies. Various degrees <strong>of</strong><br />

crookedness result from the lateral branches or buds assuming dominance over terminal<br />

shoots. Attack usually becomes less intensive as trees increase in height.<br />

In <strong>Canada</strong>, the European pine shoot moth is an important pest <strong>of</strong> pine plantations in<br />

southern Ontario, but it also exists in southern Quebec, primarily on ornamentals<br />

(Beique 1960). However, the Forest Insect and Disease Survey has not collected the<br />

moth in Quebec since 1968. Presumably the distribution within that province is unchanged<br />

but the prevalence is markedly decreased since 1968. In British Columbia the insect is<br />

well established, mostly on ornamental stock, in the Vancouver-Victoria area and to a<br />

lesser degree in the Okanagan Valley north to Kamloops (C.E. Brown 1981 personal<br />

communication). The European pine shoot moth was first reported in the maritimes at<br />

Bear River, Nova Scotia, in 1925 (Reeks et a1."1951), and has spread to cover all Nova<br />

Scotia, Prince Edward Island, and most <strong>of</strong> the southern half <strong>of</strong> New Brunswick. It has<br />

become a serious pest <strong>of</strong> Scots and red pine plantations in Prince Edward Island and<br />

Nova Scotia, where increased plantings have intensified the problem. However, there<br />

has been little spread in New Brunswick in the last decade and abundance has remained<br />

at a low level (L.P. Magasi 1981 personal communications). Since 1968 the European<br />

387


388 P. D. Symc


Rhyaciollia blloliallll (Schiff.). 3M!)<br />

pine shoot moth has remained an unimportant pest <strong>of</strong> ornamental pines in the cities <strong>of</strong> St.<br />

John's and Comer Brook in Newfoundland (A.G. Raske 1981 personal communication).<br />

The progressive distribution <strong>of</strong> the shoot moth in North America since 1968 suggests<br />

that the shipping <strong>of</strong> infested nursery stock has contributed strongly to its dispersal,<br />

resulting in a rapidly spreading but spotty distribution. This was particularly evident in<br />

the Maritime Provinces. The Forest Insect and Disease Survey has recorded the spread<br />

in that area as well as in British Columbia. Declining populations were noted in Quebec.<br />

In Ontario, the nothern limit <strong>of</strong> continuous distribution has held, as predicted (Green<br />

1962), at about the -29"C mean minimum temperature isotherm. The populations on<br />

Manitoulin and Cockburn islands in the North Channel <strong>of</strong> Lake Huron have collapsed<br />

because the host trees have outgrown susceptibility, and populations declined generally<br />

throughout the province during the 1970s. Changes in the economics and marketing <strong>of</strong><br />

red pine, at least in Ontario, have further reduced the importance <strong>of</strong> the European pine<br />

shoot moth; deformed stems may be acceptable under some circumstances.<br />

In the rest <strong>of</strong> <strong>Canada</strong>, the situation remains static. In New Brunswick it is not known<br />

why the insect has not penetrated to the northern part <strong>of</strong> the province, although a lack <strong>of</strong><br />

susceptible hosts could be an important factor.<br />

Background Although considerable work was done on various biological aspects <strong>of</strong> the European<br />

pine shoot moth before 1968 (Syme 1971), the research emphasis and the parasitoid<br />

monitoring capability changed after 1969. For example, studies on R. buolialla in<br />

Ontario were virtually confined to the effects <strong>of</strong> wild carrot, Daucus carola L.. and<br />

other flowering plants on the effectiveness <strong>of</strong> Orgi/us obscuralor (Nees), the most<br />

effective introduced parasitoid <strong>of</strong> the pest. Meanwhile, the earlier literature on the<br />

biology. behaviour, history, distribution, ecology. damage, and control were well reviewed<br />

(Pointing & Green 1962, Pointing & Miller 1967, Miller 1967, Syme 1971). Since 1973<br />

virtually no new research has been done in <strong>Canada</strong> on either the European pine shoot<br />

moth or its parasitoids.<br />

Releases and Recoveries The following brief descriptions <strong>of</strong> species <strong>of</strong> parasitoid include recoveries or significant<br />

range extensions <strong>of</strong> species released before 1968, and significant developments and<br />

findings on their possible effectiveness determined since 1968. The origin and numbers <strong>of</strong><br />

released parasitoids are summarized in Tables 97 and 98. With the exceptions <strong>of</strong> O.<br />

obscurator (Nees), Lypha dubia (Fall.), and Parasierola lIigri/emllr (Ashm.). little can<br />

be reported.<br />

Table 97 Open releases and recoveries in Ontario <strong>of</strong> parasitoids against Rllyacionill bllolimuJ (Schiff.)<br />

Species Year <strong>of</strong> release Origin No. Released Year <strong>of</strong> recovery<br />

Agathis binominata<br />

Mues. 1969 Ontario 52<br />

1970 Ontario 33 1973<br />

Lyplla dubio (FaiL) 1969 Germany 496 1970<br />

Orgilus obscurator (Nees) 1969 Germany 218<br />

1970 Germany 45<br />

1971 Germany 225<br />

1972 Germany 82 1978<br />

Parasierola nigrifemur<br />

(Ashm.) 1974 Argentina (l}5


390 P. D. Syme<br />

Table 98<br />

Eulimneria runfernur<br />

(Thorn.)<br />

(Hymenoptera:<br />

Ichneumonidae)<br />

Lypha dubla (FaD.)<br />

(Diptera: Tachinfdae)<br />

Orgilus obscurator<br />

(Nees)<br />

(Hymenoptera:<br />

8racooidae)<br />

Laboratory and field cage studies in Ontario or parasitoids against Rhyacionia<br />

buoliana (Schiff.)<br />

Species Year Origin Number<br />

Orgilus obscurator (Nees) 1969 Germany 778<br />

1970 Germany 449<br />

1971 Germany 500<br />

1972 Germany 181<br />

1973 Germany 6<br />

Parasierola nigri/emur (Ashm.) 1973 Argentina 56<br />

E. Tuft/emur was released before 1959 at Toronto, Niagra Falls, and Elmira (McGugan &<br />

Coppe11962) but has not been released since. Although it had spread significantly west<br />

<strong>of</strong> the release point by 1968 (Syme 1971), collapse <strong>of</strong> host insect populations made the<br />

parasitoid difficult to find. It was taken only once (1972) at Elmira and twice (1969,1971)<br />

near Drayton, 16 km NW <strong>of</strong> Elmira, always in small numbers. Because this parasitoid<br />

interacts with O. obscuTator. to the latter's detriment, and in view <strong>of</strong> the relatively<br />

greater importance <strong>of</strong> O. obscurator. E. ruft/emur is considered detrimental to the<br />

control effectiveness <strong>of</strong> the two parasitoids.<br />

This species was described in 1810 from Sweden (Stone et 01. 1965) and is recorded by<br />

Thompson (1951) from many species <strong>of</strong> Lepidoptera in Europe. It is a primary larval<br />

parasitoid that deposits living larvae on twigs or foliage near host larvae. The wide range<br />

<strong>of</strong> host habitats suggests that a complex <strong>of</strong> closely related species may be involved. Small<br />

numbers were released in Nova Scotia against the winter moth, Operophtera brumata<br />

(L.), but none was recovered (Embree 1971).<br />

L. dubia attacks late-instar larvae <strong>of</strong> the European pine shoot moth, and after rapid<br />

development within the host the parasitoid larva drops to the ground to pupate. It spends<br />

most <strong>of</strong>the summer and all winter in the ground as a fully developed adult, which emerges<br />

at about the time the R. buoliana begins feeding in the spring. There is a pre-oviposition<br />

period <strong>of</strong> about 4 weeks.<br />

Studies in Europe showed that there was no serious interference by this species with<br />

Orgi/us obscurator, the most effective introduced parasitoid in Ontario, and that L. dubia<br />

is one <strong>of</strong>the more effective parasitoids in northern Germany (Adlung & Sailer 1963, D.<br />

Schroeder, CISC, Delemont, Switzerland, personal communication).<br />

The parasitoid had also been released against R. buoliana in the United States in the<br />

years 1933-38 and again in 1962 and 1963, but has not been recorded (Dowden 1962).<br />

Although 90% <strong>of</strong> 1 200 puparia set out at Elmira in 1968 successfully emerged (Syme<br />

1971), only a few were recovered and only enough in 1970 to say that at least one pair <strong>of</strong><br />

flies had successfully emerged and mated! No parasitoids have been recaptured from a<br />

release at Vance Tract, Wellington County, Ontario. Thus L. dubia, despite its effectiveness<br />

in northern Germany, does not seem to be effective against the European pine shoot<br />

moth in North America.<br />

O. obscurator is an internal solitary parasitoid that attacks first-instar larvae. Its life<br />

history has been described by Juillet (1960). Early introduction <strong>of</strong> the parasitoid and<br />

movement <strong>of</strong> nursery stock infested with R. buoliana resulted in a wide distribution <strong>of</strong> O.<br />

obscurator in Ontario (McGugan & Coppel 1962). In 1968 it was present virtually


c<br />

.2<br />

5<br />

:g<br />

Cij<br />

c<br />

Rhyacion;a buoliuna (Schiff.), 391


392 P. D. Syme<br />

Temelucha inlemlptor<br />

(Grav.)<br />

(Hymenoptera:<br />

Ichneumonidae)<br />

ParageroIa nigrifemur<br />

(Ashm.)<br />

(Hymenoptera:<br />

Bethylldae)<br />

throughout the range <strong>of</strong> European pine shoot moth in Ontario and collections since that<br />

date have confirmed this (Fig. 22). It was also well established throughout Quebec, in the<br />

maritimes, and in Michigan, where it had not been released (Syme 1971).<br />

Since 1968, this parasitoid has become widespread throughout Nova Scotia and<br />

Prince Edward Island and has been found in the southeastern corner <strong>of</strong> New Brunswick<br />

(L.P. Magasi 1981 personal communication). In Newfoundland it has been recovered<br />

twice, both times at St. John's (A.G. Raske 1981 personal communication). No releases<br />

have ever been made <strong>of</strong> this parasitoid in either the maritimes or Newfoundland. Thus,<br />

the hypothesis expressed by McGugan & Coppel (1962) and Syme (1971), that the<br />

parasitoid and its host are easily dispersed by planting infested nursery stock, is clearly<br />

reinforced by the events <strong>of</strong> later years. It is also known that O. obscurator disperses well<br />

on its own, for at Cockburn Island in Lake Huron, where it had been released in 1966<br />

(Syme 1971), it was recovered 5 years later (1971) at Sand Bay, 12 km from the release<br />

point. Here it parasitized 39% <strong>of</strong> the European pine shoot moth, indicating it had been<br />

present for some years. In 1972 no R. buo[iana could be found at Sand Bay.<br />

In British Columbia, where again no intentional releases have ever been made, O.<br />

obscllrator occurs in the general area <strong>of</strong> Vancouver-Victoria (C.E. Brown 1981<br />

personal communication) (Fig. 22).<br />

In Ontario, where further studies have been made <strong>of</strong> the beneficial effects <strong>of</strong> wild<br />

carrot and other flowering plants on the longevity and increased fecundity <strong>of</strong> O. obsauator<br />

(Syme 1977), several situations were studied since 1968 that supported this technique <strong>of</strong><br />

enhancing parasitoid attack. At one small plantation in Brantford Township, Brant<br />

County, where D. carota was plentiful, O. obscllralor was released during the summer<br />

<strong>of</strong> 1972. The population was not sampled in 1973 or 1974, but by 1975 parasitism had<br />

risen to 60% and no European pine shoot moth population remained in 1976. In three<br />

other plantations replete with D. carota, parasitism rose to over 50% in 3 to 4 years with<br />

the concurrent demise <strong>of</strong> the European pine shoot moth population.<br />

It seems apparent that the presence <strong>of</strong> a suitable source <strong>of</strong> food in the form <strong>of</strong> D.<br />

carota can, under otherwise favourable conditions, increase the effectiveness <strong>of</strong> O.<br />

obscllrator to a level virtually able to eliminate the European pine shoot moth population.<br />

On the basis <strong>of</strong> these results, workers in Nova Scotia are attempting to establish wild<br />

carrot in pine plantations, but the results are not yet fully known (L.P. Magasi 1981<br />

personal communication).<br />

This primary internal larval parasitoid with cleptoparasitic habits (Syme 1971) has not<br />

been released since 1961. It persists at Elmira, Ontario, where it competes to the<br />

detriment <strong>of</strong> O. obscllralor. T. interrllptor was virtually always in a host with O. obscllrator,<br />

resulting in the death <strong>of</strong> the latter. Thus, in a field experiment designed to show the effects<br />

<strong>of</strong> D. carola, on O. obscllralor at Elmira, from 1966 to 1978, T. interrllptor parasitized up<br />

to 50% <strong>of</strong> the hosts already attacked by O. obscuralor, preventing the latter from<br />

attaining levels <strong>of</strong> parasitism over 40% necessary to cause the collapse <strong>of</strong> host populations.<br />

Thus it confounded the field trial that otherwise showed promise <strong>of</strong> demonstrating the<br />

beneficial effects <strong>of</strong> D. carota on O. obscurator.<br />

This external, multivoltine gregarious parasitoid had been found to be very effective<br />

against European pine shoot moth in Argentina (Brewer & Varas 1971). With background<br />

supplied by E.M. deBrewer, Department <strong>of</strong> Entomology, Cordoba National University,<br />

laboratory tests in <strong>Canada</strong> (Sault Ste. Marie) demonstrated that P. nigrifemur showed<br />

no tendency to select for hosts parasitized by O. obscllralor. Thus it has no cleptoparasitic<br />

tendencies (Syme 1974). Consequently, on 21 August 1974,695 adults (unsexed)<br />

were released at a site 4.5 km north <strong>of</strong> Elsinore, Amabel Township, Bruce County,


Agatbis binominata<br />

(Mues.) (Hymenoptera:<br />

Braconidae)<br />

Evaluation <strong>of</strong> Control Attempts<br />

Rhyacionia buoliana (Schiff.) 393<br />

Ontario. The trees averaged 1.4 m in height and there was an abundance <strong>of</strong> flowering<br />

plants including D. carola. In that same year, in October, two adult P. nigrifemur were<br />

found during the annual fall sample, but they were not associated with any apparent<br />

hosts. No parasitized hosts were found during the next 2 years. We can only conclude<br />

that the attempted introduction was a failure.<br />

Although this native parasitoid was released at Vance Tract, Wellington County, Ontario<br />

in 1%9 and 1970, only a single larva was recovered in 1973 and none in 1975, 76 or 77.<br />

Reasons for its failure to establish are speculative.<br />

The difficulties <strong>of</strong> evaluating the impact <strong>of</strong> biological control agents on the European pine<br />

shoot moth were discussed by Syme (1971) and can be summarized by stating that the<br />

host-pest relationship and pest damage change drastically with time. In the absence <strong>of</strong><br />

any further trials <strong>of</strong> exotic parasitoids or further behavioural studies, the well-established<br />

European species O. obscuralor seems to <strong>of</strong>fer the greatest promise for effective<br />

localized biological control in <strong>Canada</strong>. Several cases have been documented in which, in<br />

the presence <strong>of</strong> D. carota flowers. O. obscuralor has brought about complete collapse <strong>of</strong><br />

the R. buoliana infestation within a few years.<br />

As O. obscuralor is a univoltine internal parasitoid, chemical control applied to the<br />

host larval stage in the spring will remove a proportionate part <strong>of</strong> the parasitoid population.<br />

Chemical control <strong>of</strong> the early-instar needle-mining stage will have a negative impact on<br />

the parasitoid adult population, which would be actively searching for hosts. This would<br />

depress the parasitoid population disproportionately. Thus if chemical control is required<br />

within an integrated pest management scheme, the spring wandering stage <strong>of</strong> the European<br />

pine shoot moth would be the time to spray and cause the least potential damage to O.<br />

obscurator. Similarly, a systemic insecticide applied in the spring would have the least<br />

effect on O. obscuralor, but any insecticide will depress the parasitoid population to some<br />

extent.<br />

It should be pointed out here, however, that wild carrot is listed in the secondary<br />

noxious class <strong>of</strong> the <strong>Canada</strong> Seed Act <strong>of</strong> 1961 and is listed in The Weed Control Acts <strong>of</strong><br />

Manitoba, Nova Scotia, Ontario, and Quebec. It is detrimental to cattle when they<br />

ingest large quantities, and taints the milk, but Dale (1974) has pointed out that it is not a<br />

weed <strong>of</strong> cultivated fields and prime agricultural land and that there are no reports <strong>of</strong><br />

cattle grazing on it from choice. Further, it does not compete in a well-established turf<br />

and is not found where ploughing is an annual event.<br />

Wild carrot may also affect commercial production <strong>of</strong> carrot either by transmitting<br />

pests or harbouring diseases common to both cultivated and wild varieties, or by causing<br />

the production <strong>of</strong> poor seed <strong>of</strong> commercial varieties when hybridization occurs (Frankton<br />

1955). However, in <strong>Canada</strong> carrot seed is grown commercially only in British Columbia<br />

(Dale 1974).<br />

On the positive side Dale (1974) indicates that D. carola has a relatively high nutritive<br />

value and should be regarded with greater tolerance when found in pastures among<br />

plants <strong>of</strong> low nutritive value. Thus foresters concerned with the establishment <strong>of</strong> red pine<br />

might well question the status <strong>of</strong> wild carrot as a noxious weed.


394 P. D. Symc<br />

Recommendations<br />

Literature Cited<br />

The parasitoid O. obscuralor appears to be ubiquitous and the methods for manipulating<br />

D. carola are published (Syme 1976), so it remains for the system to be practised and<br />

refined. The method may in some instances be labour-intensive. However, if a plantation is<br />

established where D. carola already occurs, the plantation manager may preserve the<br />

wild carrot to favour O. obscuralor at virtually no cost.<br />

This unique method whereby the probability <strong>of</strong> control can be enhanced by manipulating the<br />

environment into which a parasitoid is introduced, thus favouring its survival and<br />

effectiveness, represents a new working principle that requires emphasis in future<br />

research as well as in methodology associated with biological control attempts.<br />

Adlung, K.G.; Sailer, R.I. (1963) Forest entomological studies in German afforestation areas, with special regard to pine shoot moth<br />

parasitoids <strong>of</strong> Rhyacionfu buoliana (Schiff.). Allgemeine Forst- und Jagdzeirung 134.229-238.<br />

Beique, R. (1960) The importance <strong>of</strong> the European pine shoot moth Rhyacionia boo/funa (Schiff.) in Quebec City and vicinity. Canadfun<br />

Entomologist 92,858-862.<br />

Brewer. M.; Varas, D. (1971) Cria masiva de Parasierola nigrifemur (Ash.), (Hym., Bethylidac) -primeras liberuciones en Calamuchita,<br />

Cordoba, Argentina. Revista Peruana de Entomologia 14,352-361.<br />

Busk, A. (1914) A destructive pine moth introduced from Europe. Journal <strong>of</strong> Economic Entomology 7,340-341.<br />

Dale, H.M. (1974) The biology <strong>of</strong> Canadian weeds. 5. Daucus carota. Canadian Journal <strong>of</strong> Plant Science 54,673-685.<br />

Dowden, P.B. (1962) Parasites and predators <strong>of</strong> forest insects liberated in the United States through 1960. US Department <strong>of</strong> AgriculTUre<br />

Handbook 226, 70 pp.<br />

Embree, D.G. (1971) OperopltJera brumata (L.), winter moth (Lepidoptera: Geometridae). In: Biological control programmes against<br />

insects and weeds in <strong>Canada</strong>, 1959-1968. Commonwealth Instirute <strong>of</strong> Biological Control Technical<br />

Communication 4,167-175.<br />

Frankton, C.(l955) Weeds <strong>of</strong> <strong>Canada</strong>. Ottawa, Ontario; Queen's Printer, 196 pp.<br />

Green, G. W. (1962) Low winter temperatures and the European pine shoot moth. Rhyacionia buoliana (SehUf.) in Ontario. Canadian<br />

Entomologist 94,314-36.<br />

Juillet, J.A. (1960) Immature stages, life histories, and behaviour <strong>of</strong> two hymenopterous parasites <strong>of</strong> the European pine shoot moth,<br />

Rhyacionia buolfuna (Schiff.) (Lepidoptera: Olethreutidae). Canadian Entomologist 92,342-346.<br />

McGugan, B.M.; Coppel, H.C. (1962) Biological control <strong>of</strong> forest insects, 1910-1958. In: A review <strong>of</strong>the biological control attempts against<br />

insects and weeds in <strong>Canada</strong>. Commonwealth Institute <strong>of</strong> Biological Control Technical Communication<br />

2,35-127.<br />

Miller, W.E. (1967) The European pine shoot moth - ecology and control in the Lake States. Forest Science Monograph 14,72 pp.<br />

Pointing, P.J.; Green, G.W. (1962) A review <strong>of</strong> the history and biology <strong>of</strong> the European pine shoot moth, Rhyacionia buoliana (Schiff.)<br />

(Lepidoptera: Olethreutidae) in Ontario. Proceedings <strong>of</strong> the <strong>Entomological</strong> <strong>Society</strong> <strong>of</strong> Ontario 92,58-69.<br />

Pointing, P.J.; Miller, W.E. (1967) European pine shoot moth. In: Important forest insects and disease <strong>of</strong> mutual concern to <strong>Canada</strong>, the<br />

United States and Mexico. Canadian Department <strong>of</strong> Forestry and Rural Development Publication<br />

1180,162-166.<br />

Reeks, W.A.; Forbes, R.S.; Cuming, F.G. (1951) Canadfun Department <strong>of</strong> Forestry and Rural Development Annual Report, Forest Insect<br />

and Disease Survey 1950, pp. 5-20.<br />

Stone, A.; Sabrosky, C.W.; Wirth, W.W.; Foote, R.H.; Coulson, J.R. (1965) A catalogue <strong>of</strong> the Diptera <strong>of</strong> America, nonh <strong>of</strong> Mexico. US<br />

Department <strong>of</strong> AgriculTUre Handbook 276, 1696 pp.<br />

Syme, P.O. (1971) Rhyacionia buoliana (Schiff.), European pinc shoot moth (Lepidoptera: Olethrcutidae). In: Biological rontrol programmes<br />

against insects and weeds in <strong>Canada</strong>, 1959-1968. Commonwealth Institute <strong>of</strong> Biological Control Technical<br />

Communication 4,194-205.<br />

Syme, P.O. (1974) Interaction between three parasitoids <strong>of</strong> the European pine shoot moth. Canadian Forestry Service Bi-monthly Research<br />

Notes 30(2),9-10.<br />

Syme, P.O. (1976) Red pine and the European pine shoot moth in Ontario. Canadian Forestry Service Information Report O-X-244, 17 pp.<br />

Syme, P.O. (1977) Observations on the longevity and fecundity <strong>of</strong> Orgilus obscurator (Hymenoptera: Braconidae) and the effects <strong>of</strong> certain<br />

foods on longevity. Canadfun Entomologist 109,995-1000.<br />

Thompson, W.R. (1951) A catalogue <strong>of</strong> the parasites and predators <strong>of</strong> insect pests. Sect. 2, Pt. 1. Hosts <strong>of</strong> the Coleoptera and Diptera.<br />

Commonwealth Institute <strong>of</strong> Biological Control, 147 pp.


396 Appendix<br />

Raske, A.G., Newfoundland Forest Research Centre, Canadian Forestry Service, Environment <strong>Canada</strong>, Bolt 6028, St. John's, Newfoundland<br />

AIC SX8.<br />

Schooley, H.O., Pctawawa National Forestry Institute, Canadian Forestry Service, Environment <strong>Canada</strong>, Chalk River. Ontario KOJ lJO.<br />

Schroeder, D., European Station, Commonwealth Institute <strong>of</strong> Biological Control. I Chemin des Grillons, CH·2800 Delemont, Switzerland.<br />

Shemanchuk, J.A., Researeh Station, Agriculture <strong>Canada</strong>, Bolt 3000, Lethbridge, Albena TlJ 4BI.<br />

Shepherd, R.F .• Pacific Forest Research Centre. Canadian Forestry Service. Environment <strong>Canada</strong>, S06 Wcst Burnside Road, Victoria,<br />

British Columbia V8Z IMS.<br />

Smirn<strong>of</strong>f, W.A.. Laurentian Forest Research Centre, Canadian Forestry Service, Environment <strong>Canada</strong>, Bolt 3800. Stc·Foy. Quebec GlV 4C7.<br />

Syme, P.D .• Great Lakes Forest Research Centre, Canadian Forestry Service, Environment <strong>Canada</strong>. Bolt 490. Sault Stc. Marie, Ontario<br />

P6A SM7.<br />

Thompson. L.S .• Researeh Station. Agriculture <strong>Canada</strong>, Bolt 1210. Charlottetown, Prince Edward Island CIA 7M8.<br />

Tumock, W.J., Research Station, Agriculture <strong>Canada</strong>, 19S Dafoe Road, Winnipeg, Manitoba R3T 2M9.<br />

TyrreD, D., Forest Pest Management Institute, Canadian Forestry Service. Environment <strong>Canada</strong>, Bolt 490. Sault Ste. Marie. Ontario P6A 5M7.<br />

Van Sickle, G.A .• Pacific Forest Researeh Centre, Canadian Forestry Service. Environment <strong>Canada</strong>. 506 West Burnside Road, Victoria,<br />

British Columbia V8Z IMS.<br />

Vanyo I.W., Maritimes Forest Researeh Centre. Canadian Forestry Service. Environment <strong>Canada</strong>. Bolt 4000. Fredericton. New Brunswick<br />

E3B SP1.<br />

Watson. A.K .• Macdonald Campus, Faculty <strong>of</strong> Agriculture. McGill University, Ste-Anne de Bellevue, Quebec H9K lCO.<br />

Whisler, H.C .• Depanmcnt <strong>of</strong> Botany, University <strong>of</strong> Washington. Seattle. Washington. USA 9819S.<br />

Wilkinson. A.T.S, Research Station, Agriculture <strong>Canada</strong>. 6660 N.W. Marine Drive, Vancouver. British Columbia V6T IX2.<br />

Wilson. G.G., Forest Pest Management Institute. Canadian Forestry Service. Environment <strong>Canada</strong>, Bolt 490. Sault SIC. Marie, Ontario P6A 5M7.<br />

Wylie, H.G., Research Station, Agriculture <strong>Canada</strong>, 19S Dafoe Road, Winnipeg, Manitoba R3T 2M9.<br />

Zebold, S.L. Depanment <strong>of</strong> Botany. University <strong>of</strong> Washington, Seattle, Washington. USA 9819S.


Index <strong>of</strong> Species<br />

Abies amabilu 229<br />

abdulghani, Telraphleps<br />

Abiesbalsamea 229,235,236.241.242,251,252,254.256.<br />

263,271.321,359<br />

Abiesfirma 267<br />

Abies/raseri 229<br />

Abies grandis 229<br />

Abies lasiocarpa 229<br />

Abies spp. 226.229.267<br />

abielu. Neodiprion<br />

abruplorius. Exenterus<br />

absinth - see Arlemisia absinlhium<br />

absinthium. Arlemisia<br />

acanthium. Onopordum<br />

acanthoides. Carduus<br />

acaulis. Silene<br />

Acernegundo 315<br />

Acersaccharum 314.349,350<br />

Acerspp. 353<br />

Aceriaacropliloni 105<br />

Achilleafilipendulina 114<br />

Achilleaplarmica 114<br />

Achromobaclerspp. 35<br />

Acleris varian a 321<br />

Acrididae 5<br />

acridophagus, Nosema<br />

acroplili. Puccinia<br />

A cropli/on repens 97, 105·110<br />

acropli/oni, Auria<br />

acroplilonica, Aulacida<br />

AClinocephalus lipulae 85<br />

Acyrthosiphon neerlandicum 160<br />

Acyllhosiphonpuum 4,7<br />

Acyrthosiphum cyparissiae 168<br />

Adalia IUleopicla 230<br />

Adalia ronina 230<br />

Adalia lelraspilola 230<br />

Adelges merkeris 234<br />

Adelges nusslini 234<br />

Adelges piceae 215,220,229·234,321<br />

adelgid - see Adelges spp.<br />

adelphocoridis, Perulenus<br />

Adelphocoris lineolalus 4,5,9·10<br />

adonidis, Emomosce/is<br />

adslriclus, Pleroslichus<br />

Aedes aegypli 24<br />

Aedes dorsalis 23<br />

Aedes jlavescens 24<br />

Aedes sierrensis 24<br />

Aedes triseriatus 24<br />

Aedes vexans 23<br />

atgypli, Aedes<br />

atreus, Monodontomerus<br />

atrogenes, Enterobacler<br />

atlhiopoides, Microclonus<br />

a/finu. Exenttrus<br />

a/finis, Urophora<br />

Agapela zoegana 129,136<br />

Agalhis binominQla 218,389,393<br />

AgQlhis cincta 287,289<br />

Agathispumila 218,281·282,283.287<br />

Agriahousei 217,218.267,268.271<br />

Agri/ushyperici 98,172.174,176<br />

Agromyza/rontella 4,11·13<br />

Agropyron crislalum 121<br />

Agropyron repens 121<br />

Agryponjla\'eolatum 218.353.354·355.356<br />

alba, Populus<br />

alba. Silene<br />

albicans. Cyzenis<br />

a/Monica. PhyllOlrela<br />

albipes. Gryocenlrus<br />

Albugo Iragopogi 112<br />

alder-see Alnus spp.<br />

aldrichi, Sarcophaga<br />

alfalfa 7.9. 10. II. 41. 45. 46<br />

alfalfa blotch leafminer - see Agromyza /rontella<br />

alfalfa mosaic virus 205<br />

alfalfa plant bug - see Ade/phocoris lineolalus<br />

alfalfa weevil-see Hypera po.flica<br />

A/nusspp. 287<br />

alpine larch - see Larix Iyafli<br />

Aisophilapomelaria 223.354<br />

aisophilae. Telenomus<br />

Allicacarduorum 98.139-140.143<br />

amabilis, Abies<br />

amabilis fir - see Abies amabilis<br />

amaranthoides. Axyris<br />

Amerboa moschata 129<br />

Ambrosiaartemisii/olia 100.111-112<br />

Ambrosia cllamissonu III<br />

Ambrosiapsilostacltya III<br />

Ambrosia lenui/olia III<br />

Ambrosia Irifida III<br />

Ambrosiaspp. III<br />

americana. Entomoscelis<br />

americana. Sorbus<br />

americana. Ulmus<br />

amiclOrius. Exenterus<br />

ampelus. Mericia<br />

amp/iala, Tingis<br />

amy/ovorus. Bacillus<br />

Anaitisp/agiala 96.172.174.176.177<br />

Anaslalus disparis 218.306-307,308.309<br />

andalusica. Hadena<br />

annual sow-thistle -see Soncllus oleraceus<br />

Anopheles spp. 225<br />

antelope 183<br />

Amltemis lillcloria 114<br />

antiqua, Delia 4<br />

antirrltinii. Gymnaelron<br />

antisypltililica, Euphorbia<br />

ants 162,313-seealsoFormicaspp.<br />

anurus. Balhyplecles<br />

Apanteles arisba 70<br />

Apanleles circuTnscriplus group 70<br />

Apanteles circuTnscriptus 70<br />

Apanteles coleophorae 218. 287<br />

Apanteles corvinus 218. 287<br />

Apanteles /umi/eranae 271<br />

Apanle/esg/omeralus 16.17<br />

Apanteles lacleicolor 305. 308<br />

Apanteles laelus 70<br />

Apanteles laeviceps 34<br />

Apanteles melanoscellus 302<br />

Apanteies me.wxamlws 21 R. 287<br />

Apanteles mililaru 34<br />

Apanteies ornigu 69. 70<br />

Apanleles pedias 70<br />

Apanteles pllllellae 15. 17. 18<br />

397


398 Index <strong>of</strong> Species<br />

Apanlt'les rubecula 15.17. 18<br />

Apanlt'les solilarius 299.301<br />

Apanlt'les xylinus 50<br />

Apanteles spp. 58.286.287.288.289.302<br />

Aphidecla. oblilerala 230<br />

Aphidencyrllls aphidivertls 173<br />

Apllidius cardui 173<br />

aphidivmlS, AphidencYrlllS<br />

Aphidoletes thompsoni 230<br />

AphisclJ/oris 97.172-173.174.176,177<br />

Aphiseuphorbiae 168<br />

Aphlhona cyparissiae 168<br />

Apillhona czwalinae 168<br />

Apillhonaf!ava 168<br />

Apillhona spp. 168<br />

apple 25.69.101.211.353<br />

Archipscerasil'oranus 260<br />

Archipsoporana 226<br />

arclicus, Sorex<br />

A renelra rufipes "emalis 34<br />

arisba. Apameles<br />

armyworm-see Mameslraconfigurata<br />

Arlemisiaabsinlilium 113-114<br />

Artemisiacana 114<br />

A rlemisia dracun,ulus 114<br />

A rlemisia longifolia 114<br />

A rtemisia vulgaris 114<br />

Artemisiaspp. 114<br />

artemisiifolia, Ambrosia<br />

artichoke plume moth - see Platyptilia cardl/idaClyla<br />

artichoke - see Cynara scolymus<br />

Artogeia rapae 4. 15-18.81<br />

arvense, Cirsium<br />

arvense, Thlaspi<br />

arvensevar. inlegrifolium, Cirsium<br />

arvensis, Convolvulus<br />

an'ensis, Sonehus<br />

arvensis s. str .• Sonehus<br />

arvensis var. glabreseens, Sonehus<br />

Asclepias syriaeQ 99-see also milkweed<br />

aspen - see Populus spp.<br />

asper, Sonehus<br />

aster yellows 205<br />

alalanlae fulveseens, Theronia<br />

Athrycia cinerea 50.54<br />

A thrycia erythroeera 53<br />

Atilryeia impressa 53<br />

allantieus, CeUlorhyneilus<br />

alralula, Leueopis<br />

atricapilano, Coehylis<br />

Queuparia,Sorbus<br />

Aulacidaacroptiloniea 105.109<br />

auricularia, Forfieula<br />

Austrian pine - see Pinus nigra<br />

Axyrisamaranliloides 121<br />

azurea, Cassida<br />

Bacillus amylovorus 211<br />

Bacilluscereus 35,313<br />

Bacillus sphaerieus 35<br />

Baeilluslhuringiensis vii,3,15,18,33,35,57.58,79,80,87,<br />

215.217,218-219,221,222,224,225,237,238-247.264.<br />

265,274,276,278,308,309.313.314-315.317.318,356.<br />

365-366.367<br />

BacillllS Ihuringiensis var. israelellsis 20.21<br />

Bacillus IhurillgiellSiHar. kurslaki 15.238<br />

bacu10virus-see viruses<br />

bakeri, Copidosoma<br />

Balauslium sp. 231<br />

ball cactus - see Mammillaria vivipara<br />

Ballia el/c/Jaris 231<br />

balsam fir-see Abies balsamea<br />

balsam fir sawny - see Neodiprion abietis<br />

balsam poplar-see Popl/lllS balsamifera<br />

balsam woolly ade1gid - see A delges piceae<br />

balsamea, Abies<br />

balsamifera, Populus<br />

Ballel,us f!aveseells 50.51.54<br />

ballksialla. Pill us<br />

bardus, Larinus<br />

barley 155<br />

basizOIllIS, PleolopllllS<br />

bassialla, BeaUl'eria<br />

balalas, Ipomoea<br />

BalhypleCles allllTllS 41<br />

BalhypleCles curculionis 41. 42<br />

BalhypleCles slenosligma 41<br />

Bayeriaeapiligena 168<br />

BeQuveriabassiana 219.223,313.344<br />

Beal/variasp. 302<br />

beech - see FagllS grandifolia<br />

begini, DiglypllllS<br />

benefaclor, Olesieampe<br />

bertha armyworm - sec ,\-Iamestra eonfigurata<br />

Bessa haTl'eyi 371.372<br />

Belulalenta 291<br />

Betula papyrifera 285, 291. 349.350<br />

Betula populi folia 287. 291<br />

Belula verrueosa 291<br />

Belulaspp. 291<br />

bieolor, Eupleetrus<br />

bieolor, Metriona<br />

bieolor, Mieroetonus<br />

biebersteinii, Centaurta<br />

biennis, Chorisloneura<br />

Bigoniehetasetipennis 39,40<br />

bimaeulatipennis, Sympiesis<br />

bindweed - see Com'olulus arvensis<br />

binominata, Agathis<br />

binuc/eatum, Nosema<br />

bipinllalllS, Cosmos<br />

birch-see Betula spp.<br />

birch casebearer - see Coleophora serratella<br />

birch leafminer - sec Fen,usa pusilla<br />

birds 85,129.187.258.313.343.377-378<br />

Bislonfiduearius 168<br />

bivillalus, Melanoplus<br />

black cottonwood - sec Populus triehoearpa<br />

black spruce - see Pieea mariana<br />

bladder campion - see Silene eucubalus<br />

blaneardella, PI,yllonoryeter<br />

Blepharipa pralensis 306<br />

Blondelia lIigripes 53<br />

blueberry 359<br />

bohemiea, Drino<br />

borealis, Campoplex<br />

borealis, LygllS<br />

Borrelinavirus - see viruses<br />

Braehymeria intermedia 306<br />

Brachypterolus puliearillS 180, 181


Brassica campestris - sec canola. rape<br />

Brassica hina - see mustard<br />

Brassica juncea - see mustard<br />

Brassicanapus 211-see alsocanola. rape<br />

brassicae, Mamestra<br />

breili, Harmonia<br />

brel'ifaciens, Clostridium<br />

broad leaved toadnax - sec Linaria dalmatica<br />

browntail moth -see Euproctis chrysorrllOea<br />

Bruce span worm - see Operopll/era bruceata<br />

bruceala, Operophtera<br />

brumala, Operopll/era<br />

Brussels sprouts 17<br />

buckthorn -see Rhamnus cathartica<br />

buckwheat 222<br />

bUCl'rus, Catcoblastis<br />

bUdworm - see Acleris I'ariana, Choristmleura spp.<br />

buffalo weed - see Ambrosia tri{tda<br />

bull thistle - see Cirsillm l'll/gare<br />

bllllalus. Geocoris<br />

blloliana, Rhyacionia<br />

cabbage 51<br />

cabbage looper - see TriclropllLria tli<br />

cabbage moth -see Mamestra brassicae<br />

cabbage worm - see A rtogeia rapae<br />

Cacloblastis bllcyrus 183·184<br />

CaclOblastis doddi 183-184<br />

cactus - see Mammillaria I·i\·ipara, Oputrtia spp.<br />

calcilrapaevar. cenraurea, Puccitlia<br />

calidllm, Calosoma<br />

California encephalitis 23<br />

califomica. Coccinella<br />

caliginosus, Harpalus<br />

Caloplrasia Illnuia 97.98. 180-182<br />

Calosoma calidum 34<br />

Calosoma sycoplranra 309<br />

calycinum, Hypericum<br />

Calyslegia spp. 155. 157<br />

Camnula pelillcida 4, 61-62<br />

campion - see Silene cucuballlS<br />

Campolelis flavicinclus 34<br />

Campolelis sp. 34<br />

Camponol/IS lrerCIIleanrlS 342<br />

Campoplegini sp. 53<br />

Campoplex borealis 218.287<br />

Campoplex spp. 218. 286. 287.288. 289<br />

Campylomma verbasci 211<br />

cana, Artemisia<br />

<strong>Canada</strong> thistle - sce Cirsium arl'(mse<br />

canadensis. Exenrerus<br />

x canadensis, PoplI/lLr<br />

cankerworm - see Alsophiia pomelaria<br />

cano1a 49<br />

capi/igena, Bayeria<br />

capsulae. DasitlellTa<br />

carabids 162.343<br />

cardui, Aplridius<br />

.carelui. Uroplrora<br />

carduidactyla. Plalyplilia<br />

carduorum. Allica<br />

CardUUSaCalllhoides 97.115-125<br />

CarduusnUlans 96.97.99. J(XI. J06.109. 115·125<br />

CarduuiS nutaniS ssp. macrolepis 116<br />

Carduus personala 117. 118<br />

('meill/u p),ctlOceplra/lls 124<br />

Carel,ll/.r tetlllifloTlu 124<br />

Carc/Illlsspp. 117.145<br />

caritlalCl. EtIlylia<br />

Carolina poplar - sec POlmlus x canadensis<br />

carolCl, DallCilS<br />

carrot 222.389.392.393.394<br />

carthami, I'uccinia<br />

CarthamlLHinclOriu.r 129. 139<br />

cascbcilrcr -sec Coleop/rora spp.<br />

Cassiela aZllrea 203.2o.t<br />

Cassiela hemisphaerica 203<br />

Cassida rubiginosa 139. 141. 145<br />

cassielea. Cllelymorp/ra<br />

Ctllhartica. Rhamnus<br />

Celyp/ra rosaceana 206.108<br />

Centaureabiebersleitlii 130.132-133<br />

Cetllaurea cineraria 129<br />

CetllaureadiffrlSa 96.97.99.106.127-137<br />

Centaureajurineafolia 133<br />

Cetllaureamacll/osa 96.97.99.106.127-137<br />

Cetllallrea tligra 129<br />

Cetllaureaxpralensis 106<br />

CetllUllTea ragusina 129<br />

Cetllallrea reperrs - see A croptilon repens<br />

Cetllaurea rhetlana 130<br />

Centallrea I'Qllfsiaca 130. 132. 133<br />

Centallreaspp. 129<br />

centallreae. P'lccinia<br />

Cephaloglyplalaricis 218.268-269<br />

CephaloglYPla mllrinanae 218.268.269<br />

Cephalosporium coccorum 233<br />

Cel}halosporillm sp. 233<br />

cerasil·oratlllS. Archips<br />

cereilileilfbeetie-see Oulemamelanopus<br />

cerealella. Sitolroga<br />

cereals 7 -see also grain<br />

cerellS. Bacillus<br />

Ceroma.riasp. 53<br />

Index <strong>of</strong> Species 399<br />

CellllrorlryneilidillS IlOrrid,lS- sec Triclrosirocalus horridus<br />

CeulOr/r ynclllls atlatllicus 200<br />

CeulOrl,yncllllslilllra 97.139.140-141.143.144.145<br />

Clraelogena sp. ciaripennis group 50<br />

ClraelOgena sp. 50<br />

chaleilOtIla, Marellanria<br />

Chamaesplreciaempiformis 98.163-165.168<br />

Chamaesplreciatenrhrediniformis 98.160.163-165.167.168<br />

Clramaesplrecia spp. 161<br />

chamissOllis. Ambrosia<br />

Clrelinielea viltig('r 183<br />

Chelymorpha cassielea 156<br />

chitletuis, Di/lIItlllls<br />

Chiritia gllllalCl 156<br />

ell loris, Aphis<br />

ClrorislOnellrtl bietltlis 248<br />

ClrorislCmeura diversana 267<br />

Clroristotlellra fumiferana 215. 217. 218. 219. 220. 221. 222.<br />

224.226.235-237.238-276.277.295.321.341.346.366<br />

ClroriSlOtlellra mllritlana 226.267<br />

ClrorislOnellra occidenralis 215. 217.219. 221. 257.277-279<br />

ClrorisICJtlellra pitlll.r 260<br />

ClrorislOm'llraspp. 226.260.267<br />

chrysanthemums 77<br />

chrysocepllala. Psylliodes<br />

Clrrysocllari.r cllSpieiogasler 69


400 Index <strong>of</strong> Species<br />

Chrysoc/laris laricinellae 69.218.281.282.283<br />

Chrysoc/laris pUllclifacies II. 12. 13<br />

Chrysoc/laris IricinclIlS 69<br />

Chrysocharis spp. 287<br />

Cllrysolina hyperici 96. 171. 173-175. 176<br />

Cllrysolinaquadrigemina 96.97.171.172.173.174.175.176<br />

Cllrysolina I'arians 98<br />

Chrysopa sp. 74<br />

chrysorrhoea, Euproclis<br />

cincla, Agalhis<br />

cinclipes. Exelasles<br />

cinclillrorax. Cirrospilus<br />

cilleraria. Celllaurea<br />

cinerea. Alhrycia<br />

cinereus cinereus. Sorex<br />

cinerosella, Euzophera<br />

cinnabar moth -see Tyria jacobaeae<br />

CirculiferlenellllS 191<br />

circumscriplllS. Apallleles<br />

CirrospilllS cinclillrorax 69<br />

Cirrospilus spp. 287<br />

Cirsiumarvense 97.98,101,102.109,139-146,195<br />

Cirsium arvense var. integrifolium 140<br />

Cirsiumdrummondii 142<br />

Cirsiumflodmanii 106<br />

Cirsium spinosum 118<br />

Cirsium vulgare 96.118.147-153<br />

Cirsiumspp. 116.117,142,145<br />

claripennis, Chaelogena<br />

c1earwinged grasshopper - see Camnula pellucida<br />

clear winged moth - see Chamaesphecia lenlhrediniformis<br />

Cleonuspiger 141.145<br />

cloacae, Enterobacler<br />

Clostridium brevifaciens 313<br />

Coccinellacalifornica 7<br />

Coccinellaseplempunclala 5.231<br />

Coccinella Irifasciala 7<br />

Coccinella undecimpunclala 7<br />

coccorum, Cephalosporium<br />

coccus, Daclylopius<br />

codling moth - see Cydia pomonella<br />

Coelomomyces psorophorae 23,24<br />

cole crops 15.18<br />

Coleophora fuscedinella 285<br />

Coleophoraklimeschiella 98.191-192<br />

Coleophoralaricella 215,218.220.222.281-284<br />

Coleophoraparthenica 98.191.192<br />

Coleophoraserralella 215,218,220,222,285-289<br />

coleophorae, Apanteles<br />

colesi, Mieroclonus<br />

coli, Escherichia<br />

Collops vi"alus 74<br />

comata. Slipa<br />

common buckthorn - see Rhamnus calhartica<br />

common mullein - see Verbascum Ihapsus<br />

common ragweed - see Ambrosia arlemisiifolia<br />

commUl'is. Meleorus<br />

Compsilura concinnala 16.301.305.306<br />

compla. Linnaemya<br />

concinnala. Compsilura<br />

confera, Leuzea<br />

configurata, Mames"a<br />

confusus, DaclylopillS<br />

congesla, Colesia<br />

conica. Sympiesis<br />

Conidiobolus sp. 261<br />

ccmquisilOr, flopleclis<br />

cOllSobrina. Emeslia<br />

Contarinia sc1,lechlendaliana 208<br />

Contarillia sonchi - see Contarinia schlechlendaliana<br />

com'o/l'uli, Er)'siphe<br />

COI'I'oll'uillS an'etlSis 155<br />

COIII'o/l'ulus sepium 156<br />

Com'oil'ulusspp. 155,157<br />

corOllala f. sp. al'ellae. Puccinia<br />

Copidoso",a bakeri 34<br />

corvillus. Apallleles<br />

Corylus spp. - see filbert<br />

cosmos III<br />

Cosmos bipimralus 111. 114<br />

COlesia cOllgesla 53<br />

COlesia glomerala 53<br />

COlesia lael'ieips 50<br />

Colesia melanoscelus 306<br />

cOllonwood - see POPUlllS spp.<br />

crane fly - see Tipula paludosa<br />

Cralaegusspp. 101,353<br />

creeping red fescue 180<br />

aislalllm. Agropyroll<br />

Cremifallia IIigrocellulala 231<br />

cricket paralysis virus 248<br />

crislala. Siphona<br />

crown rust <strong>of</strong> oats-see Puccinia coronala f.sp. al'enae<br />

Cruciferae 31.32,74<br />

allfiferae. Phyllolrela<br />

cruciferous garden crops 31<br />

cserei, Silene<br />

cucubalus. Silene<br />

Cucullia verbasci 211.212<br />

cucumber mosaic virus 205<br />

cucumbers 77, 78<br />

Culex pipiens 4. 19-21,24<br />

Culex quillquefascialus 24<br />

Culex reSlllallS 19<br />

Clliisela illomala 4.23-24<br />

culpalor, Slellic/lllellmoll<br />

cunea, Hyphan"ia<br />

Cllnealum, Nosema<br />

cllrclliiollis. Balhyplecles<br />

curly top virus 191<br />

cllSpidogaster, Clrrysoc/,aris<br />

cutworm - see Euxoa messoria<br />

cyanella, Lema<br />

Cyanus segelum 129<br />

Cyclops vemalis 23<br />

C)'diapomonella 4.25-27<br />

Cynarascolymus 106.109.129,139<br />

cyparissiae. AC)'rlhosiphum<br />

C)"parissiae, Aphlholla<br />

C)"parissias. Euphorbia<br />

cypress spurge - see Euphorbia cyparissias<br />

Cyrtacanthacridinae 62<br />

Cystiphora sonchi 97.206.207,208<br />

Cyrtoga.ftersp. 11<br />

cytoplasmic polyhedrosis virus - see viruses<br />

Cyzenisalbicans 217,218.353,354-355.356<br />

czwalinae, Aphthona<br />

Dacnusa dryas 11, 12. 13<br />

Dactylopius coccus 183


pellucida. Cyrtacanthacridinae. Melanoplus spp .•<br />

Oedopodinae<br />

great ragweed - see Ambrosia trifida<br />

greenhouse whitefly-see Trialeurodes vaporariorum<br />

Gregarina longa 85<br />

gregarines 61<br />

grey birch - see Betula populi/olia<br />

ground beetles 85<br />

ground squirrel 183<br />

grylli. Entomophthora<br />

Grypocentrus albipes 218.291.292.293·294<br />

gunata. Chirida<br />

Gymnaetronantirrhinii ISO. 181<br />

Gymnaetron tetrum 211<br />

Gypsophila repens 203<br />

Habrocytus semotus 287<br />

Habrocytuselevatus 141<br />

Habrocytus spp. 287<br />

Hadenaandalusica 204<br />

Hadena Iweago 204<br />

Hadrobunus maculosus 34<br />

hares 378<br />

Harmonia breiti 231<br />

Harpaluscaliginosus 34<br />

harveyi. Bessa<br />

hastata, Rheumaptera<br />

hay 79.204<br />

Helianthus sp. III<br />

sect. Helioscopa, Euphorbia<br />

heliothidis. Heterorhabditis<br />

hemipterus, Eupteromalus<br />

hemlock looper - see Lambdina {/Scellaria {/Scellaria<br />

herculean us, Camponotus<br />

hercyniae, Borrelinavirus<br />

hercyniae, Diprion<br />

hercyniae, Gilpinia<br />

Herpetomonas swainei 344<br />

Heterorhabditis heliothidis 220<br />

hilaris, Theronia<br />

Hirmocystis ventricosa 85<br />

Hirsutellathompsoni 261<br />

histrionanna, Dichelia<br />

Horismenus/raternus 69<br />

horridus, TrichosirocaIus<br />

honensis. Miscogaster<br />

housefly -see Musca domestica<br />

housei, Agria<br />

Howardulaphyllotretae 74<br />

Howardula sp. 73<br />

Hylemya seneciella 98, 197. 199. 200<br />

Hyleseuphorbiae 96.97.98.159.161·163.168<br />

Hyperapostica 4.41·43<br />

hyperici, Agrilus<br />

hyperici. Chrysolina<br />

Hypericumcalycinum 173<br />

Hypericum densiflorum 173<br />

Hyperieumper/oratum 96.97.98.99.171·177<br />

Hypericum rhodopaeum 173<br />

Hyphantria cunea 260<br />

iberica, Salsola<br />

impexus, Pullus<br />

imponed cabbageworm - see Anogeia rapae<br />

impressa, Athrycia<br />

ineertus, Tetrastiehus<br />

inornata, Culiseta<br />

insulare, Diadegma<br />

intermedia. Brachymeria<br />

intermedius. Diglyphus<br />

interruptor, Temelucha<br />

Ipomoea batatas - see sweet potato<br />

iridescent virus - see viruses<br />

isaea, Diglyphus<br />

Itoplectis conquisitor 81.306<br />

Itoplectis spp. 287<br />

jaceae, Larinus<br />

jaceae, Puccinia<br />

jack pine - see Pinus banksiana<br />

jack pine bud worm - see Choristoneura pinus<br />

jacobaeae, Longitarsus<br />

jacobaeae, Senecio<br />

jaeobaeae, Tyria<br />

Japanese fir-see Abiesfirma<br />

japonicus, Dicladocen/S<br />

jugoslavica, Sphenoptera<br />

julis, Tetrastiehus<br />

jurinea/olia, Centaurea<br />

kali var. tenui/olia, Salsola<br />

kasoehstaniea, Urophora<br />

kinghead - see Ambrosia trifida<br />

Klebsiella pneumoniae 35<br />

klimeschiella, Coleophora<br />

klugii, Tritneptis<br />

knapweed 95. 102-see also Acroptilon repens, Centaurea<br />

spp.<br />

knapweed root moth - see Agapeta zoegana<br />

kuvanae, Ooencyrtus<br />

lateicolor, Apanteles<br />

Lactuca sativa var. capita 114<br />

laetus, Apanteles<br />

laevieeps, Apanleles<br />

laevicips, Cotesia<br />

Lambdina {/Scellaria {!Seellaria 219. 226<br />

Lambdina {/Seellaria lugubrosa 226<br />

lapponicus, Rhorus<br />

larch - see Larix spp.<br />

larch case bearer - see Coleophora larieella<br />

large tooth aspen - see Populus grandidentata<br />

larieella, Coleophora<br />

laricina, Larix<br />

laricinellae, Chrysoeharis<br />

laricinel/um, Diadegma<br />

laricis. Cephaloglypta<br />

Laricobius eriehsonii 231<br />

Larinus bardus 105<br />

Larinusjaceaf 105<br />

Larix laricina 233.281.359.369.371.381<br />

Larix Iyallii 369<br />

Larix oecidentalis 281, 369<br />

Larix spp. 226. 369<br />

larvarum. Eulophus<br />

/arvarum, Exorista<br />

lan'arum, Fusarium<br />

lasioearpa, Abies<br />

Lathrolestes nigrieollis 218.291·294<br />

lathyris, Euphorbia<br />

Index <strong>of</strong> Species 403


404 Index <strong>of</strong> Species<br />

lawn 85<br />

leafy spurge 95, 100, 102-see also EuphorbiQ esulQ·virgata<br />

complex<br />

leatherjacket-see TipulQ spp.<br />

LebiQ viridis 140<br />

LecQnium tiliae 215<br />

lecontei, Neodiprion<br />

legume crops 45 -see also alfalfa, red clover, sainfoin, sweet<br />

clover<br />

Leiobunum villQtum 34<br />

LemQcyanella 101,142,144<br />

Lespesia sp. 50<br />

lettuce 107,109,114<br />

Leucoma SQlicis 215,220,222,299·302,305<br />

Leucopis atrQlulQ 232<br />

Leucopis n .sp. nr. melQnopus 232<br />

Leucopis obscurQ 232<br />

leucostigma, OrgyiQ<br />

LeuzeQcon/erQ 129<br />

leviventris, Meteorus<br />

LinQriQ dQlmQticQ 98, 179·182<br />

LinQriQ l'ulgQris 97,179·182<br />

linden-see Tiliaspp.<br />

lineola, Thymelicus<br />

lineolaris, Lygus<br />

lineolatus, Adelphocoris<br />

Linnaemyacompta 34<br />

Liothrips sp. 111<br />

Lissonotasp. 218,267,268,270<br />

litura, Ceutorhynchus<br />

lituratus, Exochomus<br />

LobesiaeuphorbianQ 168<br />

locustae, MalamebQ<br />

locustae, Nosema<br />

loewi, Dasineura<br />

Lombardy poplar-see Populus nigra var. italica<br />

longa, GregarinQ<br />

longi/oliQ, Artemisia<br />

Longitarsus /lavicomis 198<br />

LongitaTSusjacobaeae 96,196,198·200<br />

Lophyroplectus luteator 218,313,334,335,338,339<br />

lugubris, FormicQ<br />

luna, Patasson<br />

lunula, Calophasia<br />

luteago, Hadena<br />

luteator, Lophyroplectus<br />

luteopicta, Adalia<br />

lyalli, Larix<br />

Lygus borealis 45<br />

Lygus deserlinus 45<br />

Lygus lineolaris 45<br />

Lygus rugulipennis 45<br />

Lygus shulli 45<br />

Lygus unctuosus 45<br />

Lygus spp. 4,45·47<br />

Lymantriadispar 216,217,218,219,220,221,222,225,268,<br />

303·310<br />

LyphQdubia 217,218,389,390<br />

Lysiphlebus/aborum 173<br />

maculipes, Pnigalio<br />

maculosa, Centaurea<br />

maculosus, Hadrobunus<br />

Malacosomadisstria 215,217,219,221<br />

MalQmeba locustae 61<br />

Malus spp. - see apple<br />

MamestrQ brassicae 49,5 I, 52, 53, 54, 85<br />

Mamestra configurata 4,49·55<br />

mammals, small 258,372<br />

Mammillaria vil'iparQ 183<br />

Manduca quinquemaculata 4,57·59<br />

Manitoba maple -see Acer negundo<br />

Mantis religiosa 5<br />

maple-sec Acerspp.<br />

mQrcescens, Serratia<br />

Marchantia cha/chonta 31<br />

mQrianum, Silybum<br />

maritima, Silene<br />

MaTSsoninasonchi 206<br />

marylandensis, Sympiesis<br />

masked shrew - see Sorex cinereus cinereus<br />

Matricaria matricarioides 114<br />

mQtricarioides, Matricaria<br />

MQtrusspp. 342<br />

maura, UrophorQ<br />

maxima, Tipula<br />

mediator, Microplitis<br />

MedicQgo sativa - see alfalfa<br />

Medicago sp. 11<br />

medullana, Peloclrrista<br />

Megaselia paludosa 87<br />

Meigelliamlltabilis 31,32<br />

Meigelliasp. 32<br />

Melandrium spp. 203<br />

Melanoplusbivillatus 61<br />

Melanopluspackardii 61<br />

Melanoplussanguinipes 61<br />

MelQnoplusspp. 4,61·62<br />

melanopus, Leucopis<br />

melanopus, Oulema 4<br />

melanoscelus, Cotesia<br />

melanoscellus, Apanteles<br />

Melilotus sp. 11<br />

Melitara dentata 183<br />

mella, Exorista<br />

menziesi;, PselldotsugQ<br />

Mericiaampelus 50,51,54<br />

merkeri, Adelges<br />

mermithids 73,74<br />

MeroslOmaspp. 19,21<br />

Mesochorus dimidiatus 372-373, 377, 378<br />

Mesochorus globulQtor 384, 385<br />

Mesochorusspp. 74<br />

Mesoleius tenthredinis 218,372,373,374,378,379<br />

mesoxanthus, Apanteles<br />

messaniellQ, Phyllonorycter<br />

messoria, Euxoa 4<br />

Meteoruscommunis 34<br />

Meteorus lel'iventris 34,53<br />

Meteorus versicolor 305,308<br />

MetrionQ bicolor 156<br />

MetrionQ purpurata 156<br />

MetzneriapaucipunctellQ 97,132-133,134,135<br />

mice 129,132,162,187,378<br />

Microctonus aethiopoides 41,42<br />

Microctonus bicolor 74,75,76<br />

Microctonuscolesi 41.42<br />

Microctollus villatQe 73<br />

M;croctonus sp. Z 74,75<br />

Microplitis mediator 51,53,54


.\ticroplitis pluu!llae 16. 17<br />

M icroplitis tubereulifer 53<br />

microsporidia 31. 61. 260-261, 262-265<br />

migratory grasshopper-sec Melanoplus sunguinipes<br />

militaris. Apunteles<br />

milkweed 222<br />

.'.finoamurinata 168<br />

minUlum. Trichogramma<br />

Miscogasterhortf?flsis 11, 12, 13<br />

modestus. Podi.ws<br />

moles 343<br />

MonodontomeTiu aereus 308<br />

montium. Tipula<br />

mosehata, Amberboa<br />

mosquito - see Aedes spp .. Anopheles spp., Culex spp.,<br />

Culiseta spp.<br />

mountain ash - sec Sorbus spp.<br />

mountain-ash sawfly -see Pristiphora geniculala<br />

mountain pine beetle - see Dendroetonus ponderosae<br />

Mugho pine-see Pinusmugo<br />

mugo. Pinus<br />

mullein - see VerbasClim Ihapsus<br />

mullein leafbug-sce Campy/omma verbosci<br />

mullein shark moth -see Cucullia verbasci<br />

mullilineatum. Zagrammosoma<br />

murinanae. Cephaloglypta<br />

murinana. Choristoneura<br />

murinala. Minoa<br />

Mu.rca domeslica 4.63-64<br />

M uscina slabu/ans 34<br />

mustard 31<br />

mUlabi/is. Meigenia<br />

Nabisferus II<br />

nana. Diadegma<br />

nebulosus. Euloplllls<br />

necalrix. Vairimorpha<br />

neerlandicum. AcyrtllOsiphon<br />

negundo. Acer<br />

nematodes 220,225 - see also OD-136. mermithids<br />

Neodiprionabielis 216,217,221. 222.321-322<br />

Neodiprion lecontei 215,217.220.221.222,225,323-329<br />

Neodiprion pralli banksianae 343<br />

Neodiprionsmifer 215.217.218.220.221. 225. 331-340<br />

NeodiprionsK'ainei 215.217.218.219.221.222.341-348<br />

Neopleclanacarpocapsae 87<br />

ni. Trichop/usia<br />

Nicoliana sp. 211<br />

nigra. Centaurea<br />

nigra. Pinus<br />

nigra var. italica. Populus<br />

nigricana. Epinota<br />

nigrifemur. Parasierola<br />

nigripes. Blondelia<br />

nigrocellu/ata. Cremifania<br />

nigricollis. Lathrolestes<br />

nil·ale. Fusarium<br />

noclij7ora. Silene<br />

nodding thistle - see Carduus nUlatlS<br />

northern house mosquito - see Cu/ex pipiens<br />

Nosema acridophagu.r 6 I<br />

Nosema binucleatum 85<br />

Nosemacuneatum 61<br />

Nosemadisstriae 313.314.315.317.318<br />

Nosema fumiferanae 257.260.262-264<br />

Nosema locustae 61.62.260<br />

Nosemapyraustae 260<br />

Nosemaspp. 35.51.61,141.142<br />

Notanisomorplla sp. II<br />

muslini. Ade/ges<br />

nlllans. Carduus<br />

nUlans ssp. macro/epis. Carduus<br />

oak - see Quercus spp.<br />

oak leafmincr-see Phyllonorcyler messaniella<br />

oats 155.185.188.205<br />

Obereaerylhrocepha/a 97.165-167.168<br />

oblilerala. Aphidecla<br />

obscura. Leucopis<br />

obscurator. Orgilus<br />

obscuripes. Formica<br />

occidentalis. CllOristOlleura<br />

ocridentalis. Larix<br />

Odiellus pictus 34<br />

Oedopodinae 62<br />

<strong>of</strong>ficinale. Taraxacum<br />

o/eracea. Tipula<br />

o/eraceus, Sonchus<br />

Index <strong>of</strong> Species 405<br />

O/esicampe benefactor 218.370.372.373-375.376-377.378,<br />

379.384<br />

Olesicampe genicu/alae 218.381.382.383-385<br />

Oncope/lus /ascialus 99<br />

onion maggot-see Delia amiqua<br />

onions 29.30<br />

Onopordum acanthium 106<br />

Onopordumspp. 116.117<br />

OoencyrlUS kUl'anae 218.305.306.307-308,309<br />

Operophtera bruceata 216.217. 221. 349-351<br />

Operophtera brumata 215.217. 218.220.222, 349, 353-357.<br />

390<br />

opilio. Pha/angium<br />

oporana. Archips<br />

Opumia/ragi/is 183<br />

Opumia polyacamha 183-184<br />

Orgilusobscurator 218.389.390-392.393,394<br />

Orgi/usspp. 287<br />

Orgyia/eucostigma 216.217,219,221,222.359-361.364<br />

Orgyia pseudotsugata 216.217.219.221,222,225,363-367<br />

ornatus. Eclytus<br />

ornigis. Aponte/es<br />

Oulema melanopus 4.65-67<br />

Oxydia trychiata 233<br />

Pacific silver fir - see Abies amabi/is<br />

Packard grasshopper - see Me/anop/us packardii<br />

packardii. Melanoplus<br />

Paecilomyces [arinosus 344<br />

pallipes. Peristenus<br />

paludosa, Megaselia<br />

pa/udosa. Tipula 4<br />

sect. Paralias. Euphorbia<br />

Paranguinapicridis 97.105,106-109<br />

Parasetigena si/"estris 306<br />

Parasiero/a nigrifemur 218.389,390.392-393<br />

parasitoids 275<br />

parthenica, Co/eop/lOra<br />

Palasson luna 41.42<br />

patula. Tagete.s<br />

paucipunctella. Metzneria<br />

pea aphid - see Acyrthosiphon pis"m


406 Index <strong>of</strong> Species<br />

pear 211<br />

pecoseruis, Phryxe<br />

pedalis, Pimpla<br />

pedias, Apanleles<br />

Pellochrislamedullana 129,136<br />

pellucida, Camnula<br />

Penicillium sp. 233<br />

pentatomids 162<br />

peplus, EuphorbiD<br />

perennial sow-thistle -see Sonchus arveruis<br />

perforalum, Hypericum<br />

Perislenus adelphocoridis 9-10<br />

Perislenus digoneulis 45,46<br />

Perislenus pallipes 9,45,46<br />

Perislenus pseudopal/ipes 45<br />

Perislenus rubricollis 9·10,45<br />

Perislenus stygicus 45,46<br />

Perislenus spp. 46<br />

persimi/is, Phyloseiulus<br />

peslifer, Salsola<br />

Pelrona resinella 224<br />

Pllalangium opilio 34<br />

Phobocampe disparis 306<br />

Phryxe pecoseruis 50<br />

Phryxe vulgaris 16,50,53,81,82<br />

Phyllonorycler blancardella 4,69-71, 101<br />

Phyllonorycler messaniella 70<br />

Phyllolrela albionica 73<br />

Phyllolrela cruciferae 73,74,75<br />

Phyllolreta robusla 73<br />

Phyllolretastriolata 73,74,75,76<br />

Phyllolrela undulata 74<br />

Phyllolrekl vittata 74<br />

Phyllolretaspp. 4,5,73-76<br />

phyllotrekle, Howardula<br />

phytonomi, Entomophlhora<br />

Phyloseiulus persimilis 77-78,89<br />

Picea engelmannii 363<br />

Piceaglauca 233,243,251,252,253,254,255,256,262,263,<br />

295<br />

Picea mariDna 233,235,256<br />

Picea ruberu 243,244<br />

Picea spp. 226.235.236,267,295,321<br />

piceae, Abies<br />

piceae, Adelges<br />

picornovirus-see viruses<br />

picridis. Paranguina<br />

pictus, Odiellus<br />

piger, Cleonus<br />

pigweed -see Axyris amaranthoides<br />

Pimpla pedalis 306<br />

pincushion cactus-see Mammillaria vivipara<br />

pine-see Pinusspp.<br />

pine beetle -see Dedrocolonus ponderosae<br />

Pinus banksiana 323,325.334,338,339,342,345.387<br />

pinus, Choris/oneura<br />

Pinus mugo 387<br />

Pinus nigra 387<br />

Pinus ponderosa 363.387<br />

Pinus resinosa 323,325.326.327,334.338,339.387<br />

Pinus rigida 387<br />

Pinus strobus 387<br />

Pinus sylveslris 323,334.338.339<br />

Pinus virginiDna 387<br />

Pinus spp. 226,323,387<br />

pipieru, Clllex<br />

pisllm. AcyrtllOsipllOn<br />

pitch pine -see Pinus rigida<br />

plagia/a, Anailis<br />

plains prickly·pear cactus-see Opllnlia polyacantha<br />

plant bugs-see Lygllsspp.<br />

Plalyptiliacardllidaclyla 147<br />

Pleclocephalus 129<br />

Pleislophoraschllbergi 260,263·264.315<br />

Pleolophus basizonus 218,333,335.337,338,341,342<br />

Pleolophusspp. 342<br />

pllimarius, Dianthus<br />

plume less thistle - see Carduus acantllOides<br />

PlllIeliaxyloslella 4.15-18<br />

plluellae, Apallleles<br />

pllllellae, Microplilis<br />

pnellmoniae, Klebsiella<br />

Pnigalio jlal'ipes 69<br />

Pnigalio maclllipes 11.69<br />

P"igalio Ilroplalae 69<br />

P"igalio n.sp. 11<br />

Podislls modeslIlS 343<br />

polyacalllha, Opllntia<br />

pomerlaria, Alsophiia<br />

pomonella, Cydia<br />

ponderosa pine - see Pi"us ponderosa<br />

ponderosa, Pinus<br />

ponderosae, Dendroclo"us<br />

poplar - see Popllius spp.<br />

Popllius alba 299. 301<br />

Popllius balsamifera 299<br />

Popllius x canaderuis 299. 30 1<br />

Popllius delloides 299<br />

Populus grandidentala 299.301<br />

Populus nigra var. ilalica 299<br />

Popllius lremuloides 299.301.302.311.314.315.316.349<br />

Populus lrichocarpa 299, 302<br />

Populusspp. 299,317.353<br />

Porizon spp. - see Campoplex spp.<br />

poslica, Hypera<br />

potato 211<br />

praleruis, Blepharipa<br />

x praleruis, Centaurea<br />

pratti ba"ksianae, Neodiprion<br />

prickly-pear - see OpllTllia spp.<br />

Priopoda nigricollis - see Lalhrolestes nigricollis<br />

Prisliphoraerichsonii 215.217,218.220.222.369-380.381<br />

Prisliphorageniclliala 216.218,220.222.381-385<br />

protozoa 219.223.225. 274-see also individualspecies<br />

Pru"us spp. 353<br />

Pselldomonas j1lloresceru 35<br />

Pseudomonasspp. 35.313<br />

pseudopallipes, Perislenus<br />

Pseudotsuga menziesii 171,277.363<br />

pseudotsugae, De"droclo"us<br />

pseudosugala, Orgyia<br />

psi/oslachya, Ambrosia<br />

psorophorae, Coelomomyces<br />

Psylliodes chrysocephala 74<br />

Psylliodes pllncllliula 73,75<br />

ptarmica, Achillea<br />

Pleromalus puparllm 16<br />

Pleroslichus adslriclUS 31<br />

publicarius, Brachyplerus<br />

Puccinia acroplili 105.109


Puccinia calcilrapae Vilr. centaureae 129<br />

Pllct"illiacarlhami 129,130<br />

Pllceilliacentaurrae 129,130<br />

Pllceillia coronaUl f. sp. al'enae IllS<br />

Pucciniajacear 129.130.136<br />

Puccilliaplmcli/ormis 141,145<br />

pulcherrima, Euphorbia<br />

pulchripes, Diglyplllls<br />

pulrx, Daphnia<br />

PUI/IIS impexus 232<br />

pumila, Agalhis<br />

pumilio, ScymmlS<br />

pUllcli/acies, Chr),socharis<br />

puncli/ormis, Pllccillia<br />

pllnellliala, Psylliodes<br />

pl/parum, PteromalllS<br />

purpurata, Metriona<br />

plIsilla, FenllSa<br />

pycnocephalllS, CardllllS<br />

qlladri/asciata, UropllOra<br />

qllCldrigemina, Chry.wlilla<br />

qlladriplISlIIlata, lVillthemia<br />

QuercllS garryalla 353<br />

QuercllS rubra 353<br />

QuercllS spp. 353<br />

quillqlle/ascialllS, Culex<br />

qllillquemaclI/Clla, Malldllca 4<br />

radicans, ZoophtllOra<br />

ragllSilla, Centaurea<br />

ragweed -see Ambrosia spp.<br />

ragwort -see Smecio jacobaea<br />

rain bilrrel mosquito-see Culex pipiens<br />

raoi, Tetraph/eps<br />

rapae, Artogeia<br />

rape 31,49,73,74,75,205<br />

rapeseed - see rape<br />

red clover 9<br />

red oak -see QllefL'llS rubra<br />

red pine - see PimlS rfsillosa<br />

red spruce - see Piefa rubens<br />

red turnip beetle - see Elltomoscflis americana<br />

red wood ants - see Fannica spp.<br />

redheaded pine sawny - see Neodiprion lecontei<br />

rdigiosa, Mantis<br />

repens, Acroptil()fI<br />

repellS, Agropyroll<br />

repens, Centaurea<br />

repens, Gypsophi/a<br />

resillel/a, Petrona<br />

resillosa. PillllS<br />

restuans. Culex<br />

RhClmmlS cathartica 185-189<br />

rhellalla. Centaurea<br />

Rheumaptera hastata 308<br />

RhillocylillSconicllS 96,97.100.115.116.117.119·124.125<br />

rlrodopaellm. Hypaicum<br />

RhorllS lapponicllS 383<br />

RllOfllSsp.no3 381.383.384<br />

RllOrllSsp. 218<br />

Rhyaeioniabuoliana 215. 217.211l. 220. 222. 224.387-394<br />

RhynchaenllS distans 105<br />

rickellsia 220<br />

rigida. PinllS<br />

ritro. Echinops<br />

robin 187<br />

robllSla. Phyl/otrelCl<br />

Rogas Irislis 81<br />

roll ilia. Adalia<br />

rosaceana. Ce/)'pha<br />

roses 77<br />

rubecu/a. Apantdes<br />

rubens, Picea<br />

rubiginosa. Cassida<br />

rubra. QllercllS<br />

rubricollis, PerislmllS<br />

rll/a consobrina. Vespu/a<br />

rllfa. Galeruca<br />

ru{i/emllr. EII/imlleriii<br />

rllfipennis. DI'IIdroclolluS<br />

ru{ipes l'ernalis, Armelra<br />

rufopicla, Winthemia<br />

ruglllipennis. LygllS<br />

Russian knapweed - see Acropli/on repens<br />

Russian pigweed - see Axyris amaralllhoides<br />

Russian thistle - see Salsola pesli/er<br />

rye ISS<br />

saccharum, Acer<br />

sainfoin 9<br />

SI. John 's-wort - see Hypericum per/oralum<br />

SI. Louis encephalitis 19,20,21<br />

salicis, Leucoma<br />

Salixspp. 299,301. 353<br />

Salsola iberica - see Salso/a pesli/er<br />

Salsola kali var. tenui/olia - see Salso/a peslifer<br />

Salsolapesti/a 91l,191·193<br />

sUlrgllillipes, MelalloplllS<br />

seplempullclata, Coccillel/a 6<br />

Sarcophaga aldrichi 306.313<br />

satin moth - Leucoma salicis<br />

salim var. capila, Laclllca<br />

sawny-see Neodiprionspp., Gilpinia hercyniae<br />

SeambllS spp. 2117<br />

sehlechtendaliana, Conlarinia<br />

scill/bergi, P/eisltJphora<br />

Sclerolinasc/eroliorum 129.136<br />

scleroliorum, Sc/erolina<br />

sco/ymllS, Cynara<br />

Scolosia velu/ala 187, 188<br />

Scots pine - see PimlS sy/veslris<br />

ScymnllS pumilio 233<br />

segelum, CyanllS<br />

sfguierana, Euphorbia<br />

semOIIlS. HabrocYl1IS<br />

sl'IIeciella. Hylemya<br />

Smecio jacobaea 96.98,99.195·201<br />

sepillm, Convoil'ul,lS<br />

srptempullctata. Coccillella<br />

SeplOria soncl!i·arwII.fis 206<br />

Seploria sonchi/olia 206<br />

sfricellS. SpermophagllS<br />

sericeicornis, Sympiesis<br />

serratella, ColeopllOra<br />

Serratia marcescetlS 313<br />

sali/a. Neodipri()fI<br />

setipennis, Bigonichela<br />

short ragweed - see Ambrosia artemisiifolia<br />

slrlllli, LygllS<br />

Index <strong>of</strong> Species 407


408 Index <strong>of</strong> Species<br />

sibericus. Dendrolimu.r<br />

sierrensis. Aedes<br />

Silene a(aulis 203<br />

Silene alba 203<br />

Silene cserei 203<br />

Silenecllcubalus 203-204<br />

Silene glauca 203<br />

Silene maritima 203<br />

Silene noctif/ora 203<br />

Silene spp. 203.204<br />

silver poplar - see Popllius alba<br />

silvestris. Parasetigena<br />

Silybummarianum 100<br />

Silybum spp. 166. 117. 145<br />

simi/is. Diprion<br />

Siphona cristala 53<br />

Siphonaf/avifrons 53<br />

Siphonageniculala 85.86,87-88<br />

Silotroga cerealella 270<br />

skipper - see Thymelicus lineola<br />

Solidago giganlea 99<br />

So/idagospp. 160,195<br />

so/ilarius, Apanteles<br />

solslilia/is, Urophora<br />

sonchi, Contarinia<br />

sonchi, Cysliphora<br />

sonchi, Marssonina<br />

sonchi-arvensis, Seploria<br />

sonchif<strong>of</strong>ia, Seploria<br />

Sonchus arvensis 97,98,205-209<br />

Sonchusarvensiss.str. 205<br />

Sonchus arvensis var. g/abrescens 205<br />

Sonchus asper 205-209<br />

Sonchuso/eraceus 205-209<br />

Sonchus lenerrimus 206<br />

Sonchusspp. 206<br />

SorbusamericaM 381<br />

Sorbusaucuparia 381<br />

Sorex arClicus 376<br />

Sorex cinereus cinereus 295,372, 373, 376, 371<br />

sorghum 155<br />

Sorospore/la uvella 35<br />

Spalangia endius 63<br />

spanworm - see Operophlera bruceala<br />

Spermophagus sericeus 155<br />

sphaericus, Bacillus<br />

sphaerosperma, Entomophlhora<br />

Sphenop/erajugos/al'ica 97,129,133,134,135<br />

spider mite -see Te/ranychus urlicae<br />

spiders 162<br />

spini, Thecla<br />

spinosu.m, Cirsiu.m<br />

spiny annual sow-thistle - see Sonchus asper<br />

spotted knapweed - see Centaurea macu/osa<br />

spotted knapweed root moth - see Agapela zoegana<br />

spotted tentiform leafminer - see Phyllonorycler blancardella<br />

spruce - see Picea spp.<br />

spruce budworm - see Chorisloneura fumiferana<br />

spurge - see Euphorbia spp.<br />

spurge hawkmoth-see Hyleseuphorbiae<br />

slabu/ans, Muscina<br />

starlings 34,85<br />

Slenichneu.mon cu/palor 53<br />

SlenichneumonsculellalOr-see Syspasis sculellalor<br />

stenosligma, Balhyp/ectus<br />

stinkweed - see Thlaspi arl'(!nse<br />

Stipacoma/a 121<br />

strawberries 7<br />

SlreplOCOCC/lS faecalis 35<br />

strio/ata, Phyllotreta<br />

strobus, PilllLr<br />

stygicus, Perislenus<br />

stylata, UropllOra<br />

subalpine fir-see Abies lasiocarpa<br />

subnodicomi.r, Tipula<br />

sublilicomis, Diadromus<br />

sugar beet 191<br />

sugar maple - see Acer saccharum<br />

sunflower III<br />

swuralis, Entomoscelis<br />

swuralis. Eutanyacra<br />

Swaine jack pine sawfly - see Neodiprion swainei<br />

swainei, Herpelomonas<br />

swainei. Neodiprion<br />

sweet clover 9<br />

sweet potato 157<br />

sycophanta, Calosoma<br />

sylveslris, Pillus<br />

Sympiesis bimaculatipennis 69<br />

Sympiesis conica 69<br />

Sympiesis marylandensis 69<br />

Sympiesi.r sericeicornis 69. 70<br />

syriaca, Asclepias<br />

Syspasis scU/ellator 81. 82, 83<br />

Tachina lan'arum - see Exorista larvarum<br />

Tagetespalula 114<br />

tamarack -sec Larix laridna<br />

Tanacelum I'ulgare 114<br />

tansy ragwort -sec Senecio jacobaea<br />

Taraxacum <strong>of</strong>fidnale 206<br />

Telenomusalsophi/ae 223<br />

Telenomussp. 306<br />

Temelucha interruplor 392<br />

tene/lus, Circulifer<br />

lenerrimus, Sonchus<br />

lenthrediniformis, Chamaesphecia<br />

lenthredinis, Mesoleius<br />

lenuif/oru.r, Carduus<br />

tenuifolia, Ambrosia<br />

Tephrilis dilacerala 98.206.207.208<br />

Telracampe diprioni - see Dipriocampe diprioni<br />

Telranychus ur/icae 4,77-18<br />

Telraphleps abdulghalli 233<br />

Telrapltieps raoi 233<br />

letraspilota, Adalia<br />

Tetrastichus incerlUS 41.42<br />

Tetrasticllus julis 65-67<br />

Telraslichus sp. 69<br />

Ie/rum, Gymnaelron<br />

Ihapsu.r, Verbascum<br />

Theclaspini 181,188<br />

Thelohania sp. 260,344<br />

Therion giganteum 53<br />

Theronia alalanlae fulvescens 306<br />

Theronia hi/aris 306<br />

thistle - see Carduus spp., Cirsium spp .• Salsola pes/ifer,<br />

Sonchus spp.<br />

Thalspi arvense 121<br />

thomp.roni, Aphidoletes


410 Index <strong>of</strong> Species<br />

Malacosoma disstria 221,313,314<br />

Mamestra configurata 51<br />

Neodiprion abietis 221,321-322<br />

Neodiprion lecontei 215.220,221,222-223,225,323-328<br />

Neodiprion sertifer 221,331, 334, 337-338, 339<br />

Neodiprion swainei 221,341,343-345,346<br />

neodiprionid sawflies 219<br />

Operophlera bruceata 221,350-351<br />

Operophlera brumala 356<br />

Orgyia leucos/igma 219,221,222-223,359-361<br />

Orgyia pseudotsugata 219, 221, 222, 225, 258, 363-365,<br />

365-367<br />

Thylmelicus lineola 79<br />

Trichoplusia ni 16, 18<br />

picornovirus <strong>of</strong> Melanoplus bivittatus 61<br />

tobacco streak virus 205<br />

villata, Tipula<br />

villalae, Microclonus<br />

villatum, Leiobunum<br />

villatus, Col/ops<br />

vitliger, Chelinidea<br />

vivipara, Mammillaria<br />

vole 378<br />

vulgare, Cirsium<br />

vulgare, Tanacetum<br />

vulgaris, Anemisia<br />

vulgaris, Linaria<br />

vulgaris, Phryxe<br />

wasps 162<br />

websleri, Diglyphus<br />

western equine encephalomyelitis 23<br />

western larch -see Larix occidentalis<br />

western spruce bud worm -see Choristoneura occidentalis<br />

wheat 139, 155,205<br />

white birch - see Betu/Q papyrifera<br />

white elm - see Ulmus americana<br />

white spruce -see Picea glauco<br />

whitefly - see Trialeurodes vaporariorum<br />

whitemarked tussock moth - see Orgyia leucostigma<br />

wild carrot - see carrot<br />

willow -see Salix spp.<br />

winter moth -see Operophtera brumata<br />

Winthemia deilephilae 34<br />

Winthemia quadripustulata 50,53<br />

Winthemia rufopicla 34,50<br />

xylinus, Apanteles<br />

xylostel/a, Plutel/a<br />

yellow toadflax - see Linaria vulgaris<br />

Zagrammosoma multilineatum 69<br />

zea, Heliothis<br />

Zeiraphera diniana 3TI<br />

Zeiraphera spp. 226<br />

Zeuxidiplosis giardi 98<br />

zoegana, Agapeta<br />

Zoophthora radicans 79,261,262<br />

Zygogramma tonuosa 111

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!