19.02.2013 Views

Conservation Biology of Lycaenidae (Butterflies) - IUCN

Conservation Biology of Lycaenidae (Butterflies) - IUCN

Conservation Biology of Lycaenidae (Butterflies) - IUCN

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

The <strong>IUCN</strong> Species Survival Commission<br />

<strong>Conservation</strong> <strong>Biology</strong> <strong>of</strong><br />

<strong>Lycaenidae</strong><br />

(<strong>Butterflies</strong>)<br />

Edited by T.R. New<br />

Occasional Paper <strong>of</strong> the <strong>IUCN</strong> Species Survival Commission No.8<br />

<strong>IUCN</strong><br />

The World <strong>Conservation</strong> Union


<strong>Conservation</strong> <strong>Biology</strong> <strong>of</strong> <strong>Lycaenidae</strong> was made possible through the generous support <strong>of</strong>:<br />

Chicago Zoological Society<br />

DEJA, Inc.<br />

<strong>IUCN</strong>/SSC Peter Scott Action Plan Fund (Sultanate <strong>of</strong> Oman)<br />

National Wildlife Federation<br />

World Wide Fund For Nature<br />

© 1993 International Union for <strong>Conservation</strong> <strong>of</strong> Nature and Natural Resources.<br />

Reproduction <strong>of</strong> this publication for educational and other non-commercial purposes is authorized<br />

without permission from the copyright holder, provided the source is cited and the copyright holder<br />

receives a copy <strong>of</strong> the reproduced material.<br />

Reproduction for resale or other commercial purposes is prohibited without prior permission <strong>of</strong> the<br />

copyright holder.<br />

ISBN 2-8317-0159-7<br />

Published by <strong>IUCN</strong>, Gland, Switzerland.<br />

Camera-ready copy by The Nature <strong>Conservation</strong> Bureau Limited,<br />

36 Kingfisher Court, Hambridge Road, Newbury RG14 5SJ, U.K.<br />

Printed by Information Press, Oxford, U.K.<br />

Cover Photo: Karner Blue, Lycaeides melissa samuelis (Photo: Richard A. Arnold).


The <strong>IUCN</strong> Species Survival Commission<br />

<strong>Conservation</strong> <strong>Biology</strong> <strong>of</strong><br />

<strong>Lycaenidae</strong><br />

(<strong>Butterflies</strong>)<br />

Edited by T.R. New


Preface iv<br />

Acknowledgment v<br />

List <strong>of</strong> Contributors<br />

PART 1. INTRODCTION<br />

Introduction to the <strong>Biology</strong> and <strong>Conservation</strong> <strong>of</strong> the<br />

<strong>Lycaenidae</strong><br />

T.R. NEW 1<br />

PART 2. REGIONAL ASSESSMENTS<br />

Introductory Comment 22<br />

<strong>Conservation</strong> biology <strong>of</strong> <strong>Lycaenidae</strong>: a European overview<br />

M.L. MUNGUIRA, J. MARTIN & E. BALLETTO 23<br />

Overview <strong>of</strong> problems in Japan<br />

T. HlROWATARI 35<br />

<strong>Conservation</strong> <strong>of</strong> North American lycaenids- an overview<br />

J. HALL CUSHMAN & D.D. MURPHY 37<br />

Neotropical <strong>Lycaenidae</strong>: an overview<br />

K.S. BROWN, JR. 45<br />

Threatened <strong>Lycaenidae</strong> <strong>of</strong> South Africa<br />

M.J. SAMWAYS 62<br />

Australian <strong>Lycaenidae</strong>: conservation concerns<br />

T.R. NEW 70<br />

PART 3. ACCOUNTS OF PARTICULAR TAXA OR<br />

COMMUNITIES<br />

Introductory comment 77<br />

Europe<br />

The mariposa del Puerto del Lobo, Agriades zullichi<br />

M.L. MUNGUIRA & J. MARTIN 78<br />

The Large Copper, Lycaena dispar<br />

E. DUFFEY 81<br />

The Adonis Blue, Lysandra bellargus 83<br />

Large Blues, Maculinea spp 85<br />

Polyommatus humedasae<br />

E. BALLETTO 88<br />

Polyommatus galloi<br />

E. BALLETTO 90<br />

The Sierra Nevada Blue, Polyommatus golgus<br />

M.L. MUNGUIRA & J. MARTIN 92<br />

Tomares ballus<br />

H.A. DESCIMON 95<br />

The Silver-studded Blue, Plebejus argus<br />

CD. THOMAS 97<br />

The Zephyr Blue, Plebejus pylaon<br />

M.L. MUNGUIRA & J. MARTIN 100<br />

Contents<br />

vi<br />

ii<br />

The Pannonian Zephyr Blue, Plebejus sephirus<br />

kovacsi<br />

Z. BÁLINT 103<br />

The threatened lycaenids <strong>of</strong> the Carpathian Basin,<br />

east-central Europe<br />

Z. BÁLINT 105<br />

Aricia macedonica isskeutzi 105<br />

Aricia eumedon 105<br />

Cupido osiris 106<br />

Jolana iolas 106<br />

Maculinea alcon 106<br />

Maculinea nausithous 107<br />

Maculinea sevastos limitanea 107<br />

Plebejus (Lycaeides) idas 107<br />

Polyommatus (Agrodiaetus) admetus 108<br />

Polyommatus (Agrodiaetus) damon 108<br />

Polyommatus (Vacciniina) optilete 108<br />

Polyommatus eroides 109<br />

Polyommatus dorylas magna 109<br />

Lycaena helle 109<br />

Lycaena tityrus argentifex 110<br />

Pseudophilotes bavius hungaricus 110<br />

Tomares nogelii dobrogensis 110<br />

Japan<br />

Coreana raphaelis<br />

T.HIROWATARI 112<br />

Shijimiaeoides divinus<br />

T. T.HIROWATARI 114<br />

North America<br />

Lange's Metalmark, Apodemia mormo langei<br />

J.A. POWELL & M.W. PARKER 116<br />

The Hermes Copper, Lycaena hermes<br />

D.K. FAULKNER & J.W. BROWN 120<br />

Thome'sHairstreak, Mitoura thornei<br />

J.W. BROWN 122<br />

Sweadner's Hairstreak, Mitoura gryneus sweadneri<br />

T.C. EMMEL 124<br />

Bartram' s Hairstreak, Strymon acis bartrami<br />

T.C. EMMEL & M.C. MINNO 126<br />

The Avalon Hairstreak, Strymon avalona<br />

T.C. EMMEL & J.F. EMMEL 128<br />

TheAtala, Eumaeus atala florida<br />

T.C. EMMEL & M.C. MINNO 129<br />

Smith's Blue, Euphilotes enoptes smithi<br />

T.C. EMMEL & J.F. EMMEL 131<br />

The El Segundo Blue, Euphilotes bernardino allyni<br />

R.H.T. MATTONI 133


The Palos Verdes Blue, Glaucopsyche lygdamus<br />

palosverdesensis<br />

R.H.T. MATTONI 135<br />

The Xerces Blue, Glaucopsyche xerces<br />

T.C. EMMEL & J.F. EMMEL 137<br />

The Mission Blue, Plebejus icarioides missionensis<br />

J.H. CUSHMAN 139<br />

The San Bruno Elfin, Incisalia mossii bayensis<br />

S. B. WEISS 141<br />

The Lotis Blue, Lycaenides idas lotis<br />

R.A. ARNOLD 143<br />

Additional taxa <strong>of</strong> concern in southern California<br />

R.H.T. MATTONI 145<br />

Philotes sonorensis extinctis 145<br />

Satyrium auretorum fumosum 145<br />

Plebejus saepiolus ssp 145<br />

South America<br />

Selected Neotropical species<br />

K.S. BROWN, JR. 146<br />

Styx infernalis 146<br />

Nirodia belphegor 146<br />

Joiceya praeclarus 147<br />

Areas spp 148<br />

Arawacus aethesa 149<br />

Cyanophrys bertha 149<br />

Neotropical <strong>Lycaenidae</strong> endemic to high elevations in<br />

south eastern Brazil<br />

K.S. BROWN, JR. 150<br />

Riodininae: Amazonian genera with most species very<br />

rare or local<br />

K.S. BROWN, JR. 151<br />

iii<br />

Theclinae endemic to the Cerrado vegetation (central<br />

Brazil)<br />

K.S. BROWN, JR. 152<br />

Neotropical Riodininae endemic to the Chocó region<br />

<strong>of</strong> western Colombia<br />

C.J. CALLAGHAN 153<br />

South Africa<br />

Aloeides dentatis dentatis and A. d. maseruna<br />

S.F. HENNING, G.A. HENNING & M.J. SAMWAYS.........154<br />

Erikssonia acraeina<br />

S.F. HENNING, G.A. HENNING & M.J. SAMWAYS.........156<br />

Alaena margaritacea<br />

S.F. HENNING, G.A. HENNING & M.J. SAMWAYS.........158<br />

Orachrysops (Lepidochrysops) ariadne<br />

S.F. HENNING, G.A. HENNING & M.J. SAMWAYS.........159<br />

Australia<br />

Hypochrysops spp.<br />

D.P.A. SANDS....................................................................160<br />

Illidge's Ant- Blue, Acrodipsas illidgei<br />

P.R. SAMSON..............................................................................166<br />

The Eltham Copper, Paralucia pyrodiscus lucida<br />

T.R. NEW...............................................................................166<br />

The Bathurst Copper, Paralucia spinifera<br />

E.M. DEXTER & R.L. KITCHING..........................................168<br />

The Australian Hairstreak, Pseudalmenus chlorinda<br />

G.B. PRINCE................................................................................171<br />

APPENDIX 1<br />

<strong>IUCN</strong> Red Data Book Categories 173


<strong>Butterflies</strong> have long been accorded 'special' status by many<br />

people who do not like insects. Ever since the ancient Greeks<br />

employed the same word (psyche) for 'soul' and 'butterfly', an<br />

aura <strong>of</strong> spiritual or aesthetic appreciation has enhanced their<br />

generally charismatic popularity, so that people to whom 'the<br />

only good insect is a dead insect' accept readily that butterflies<br />

merit conservation (New 1991). Perhaps the most widely<br />

appreciated butterflies are the swallowtails, birdwings and their<br />

allies (Papilionidae), which have recently received substantial<br />

conservation impetus through production <strong>of</strong> a global survey<br />

(Collins and Morris 1985) and a Swallowtail Action Plan (New<br />

and Collins 1991) based on this, and produced under the<br />

auspices <strong>of</strong> the <strong>IUCN</strong> Species Survival Commission's<br />

Lepidoptera Specialist Group.<br />

Many other butterflies are much less well known than the<br />

swallowtails, and such a comprehensive appraisal <strong>of</strong> them<br />

would be difficult or impossible to achieve. But, especially in<br />

north temperate regions <strong>of</strong> the world, very substantial<br />

conservation effort has been directed to members <strong>of</strong> the largest<br />

family <strong>of</strong> butterflies, the <strong>Lycaenidae</strong> – the blues, coppers,<br />

hairstreaks, metalmarks and related forms. In 1989, I suggested<br />

that the Lepidoptera Specialist Group should seek to complement<br />

the swallowtail studies by an appraisal <strong>of</strong> the <strong>Lycaenidae</strong>, to<br />

'round out' the emerging picture <strong>of</strong> butterfly conservation by<br />

gathering together some <strong>of</strong> the data on this family. <strong>Lycaenidae</strong><br />

exemplify a wide spectrum <strong>of</strong> concerns: populations are <strong>of</strong>ten<br />

extremely localised with colonies occupying a few hectares or<br />

less; many are associated with early successional stages <strong>of</strong><br />

vegetation in grasslands or herb associations; and many<br />

participate in subtle ecological associations with ants or (more<br />

rarely) Homoptera.<br />

Preface<br />

iv<br />

A number <strong>of</strong> species or subspecies <strong>of</strong> the <strong>Lycaenidae</strong> have<br />

been the targets <strong>of</strong> major conservation campaigns which have<br />

been vitally important in raising public awareness <strong>of</strong> insect<br />

conservation in areas where swallowtails are scarce, and it is no<br />

exaggeration to claim that they have been the most important<br />

butterfly family in fostering conservation concern in temperate<br />

regions.<br />

This is not a Red Data Book but, rather, an introduction to<br />

the conservation biology <strong>of</strong> the <strong>Lycaenidae</strong>. It emphasises the<br />

very different knowledge base available for lycaenids compared<br />

with the swallowtails, and should be a salutary warning against<br />

feelings <strong>of</strong> complacency that butterflies are well known!<br />

It draws on the expertise <strong>of</strong> many experienced practitioners<br />

<strong>of</strong> butterfly conservation, and on the large literature <strong>of</strong> lycaenid<br />

biology. The account consists <strong>of</strong> three sections. The first is a<br />

brief general introduction to the <strong>Lycaenidae</strong> and their place in<br />

butterfly conservation. The second is a series <strong>of</strong> regional<br />

overviews <strong>of</strong> lycaenids for several parts <strong>of</strong> the world where<br />

interest and knowledge has been sufficient to prepare such an<br />

essay; and the third is a series <strong>of</strong> selected case-histories or<br />

species accounts which range from the well known to the novel.<br />

This section is not in any sense encyclopedic, but provides a<br />

tentative basis for direction and for future synthesis and<br />

development <strong>of</strong> more general conservation strategies.<br />

References<br />

COLLINS, N.M. and MORRIS, M.G. 1985. Threatened Swallowtail <strong>Butterflies</strong><br />

<strong>of</strong> the World. The <strong>IUCN</strong> Red Data Book, <strong>IUCN</strong>, Gland and Cambridge.<br />

NEW,T.R. 1991. Butterfly <strong>Conservation</strong>. Oxford University Press, Melbourne.<br />

NEW, T.R. and COLLINS, N.M. 1991. Swallowtail <strong>Butterflies</strong>. An Action<br />

Plan for their <strong>Conservation</strong>. <strong>IUCN</strong>, Gland.


The task <strong>of</strong> preparing this volume has been a pleasant one,<br />

leading me to appreciate very deeply the spirit <strong>of</strong> cooperation<br />

engendered by a common concern for the well being <strong>of</strong><br />

butterflies. Most authors approached to contribute to the<br />

compilation agreed promptly and willingly, and I am very<br />

grateful for their participation and encouragement. Other people<br />

responded with constructive advice and suggestions in response<br />

to my queries, and most members <strong>of</strong> the Lepidoptera Specialist<br />

Group in 1989 and 1990 suggested possible species or candidate<br />

authors. Dr. Simon N. Stuart and his dynamic team, including<br />

Acknowledgements<br />

V<br />

Dr. Mariano Gimenez-Dixon and Ms Linette Humphrey, in the<br />

SSC Office in Gland, have been continually encouraging and<br />

supportive, and their rapid responses to my numerous questions<br />

have been appreciated greatly. Editorial advice from Dr.<br />

Alexandra Hails has also been welcomed.<br />

Transformation <strong>of</strong> a miscellany <strong>of</strong> scripts into a relatively<br />

harmonic and consistent typescript was accomplished patiently<br />

by Mrs. Tracey Carpenter at La Trobe University; she also drew<br />

a number <strong>of</strong> the figures. The photographers are acknowledged<br />

individually in the photograph legends.


Dr. R.A. Arnold<br />

Entomological Consulting Services Limited,<br />

104 Mountain View Court,<br />

Pleasant Hill,<br />

California 94523,<br />

U.S.A.<br />

Dr. Z. Bálint<br />

Zoological Department,<br />

Hungarian Natural History Museum,<br />

Baross utca 13,<br />

Budapest, H-1008,<br />

Hungary<br />

Pr<strong>of</strong>. E. Balletto<br />

Dipàrtimento di Biologia Animale,<br />

Università di Torino,<br />

V. Accademia Albertina 17,<br />

Torino,<br />

Italy -10123<br />

Dr. J.W. Brown<br />

Entomology Department,<br />

San Diego Natural History Museum,<br />

P.O. Box 1390,<br />

San Diego,<br />

California 92112,<br />

U.S.A.<br />

Dr. K.S. Brown, Jr.<br />

Departmento de Zoologia,<br />

Instituto de Biologia,<br />

Universidade Estadual de Campinas,<br />

C.P. 6109. Campinas,<br />

Sao Paulo, 13. 081.<br />

Brazil<br />

Dr. C.J. Callaghan<br />

Louis Berger International Inc.,<br />

100 Halsted Street,<br />

P.O. Box 270,<br />

East Orange, NJ 07019,<br />

U.S.A.<br />

Dr. J. Hall Cushman<br />

Center for <strong>Conservation</strong> <strong>Biology</strong>,<br />

Department <strong>of</strong> Biological Sciences,<br />

Stanford University,<br />

Stanford, California 94305,<br />

U.S.A.<br />

List <strong>of</strong> contributors<br />

vi<br />

Dr. H.A. Descimon<br />

Laboratoire de Systématique évolutive,<br />

Université de Provence,<br />

3 place Victor Hugo,<br />

13331 Marseille Cedex 3,<br />

France<br />

Ms E.M. Dexter<br />

School <strong>of</strong> Australian Environmental Studies,<br />

Griffith University,<br />

Nathan,<br />

Queensland 4111,<br />

Australia<br />

Dr. E. Duffey<br />

Cergne House,<br />

Church Street,<br />

Wadenhoe,<br />

Peterborough PE8 5ST,<br />

U.K.<br />

Dr. J.F. Emmel<br />

26500 Rim Road,<br />

Hemet,<br />

California 92544,<br />

U.S.A.<br />

Dr. T.C. Emmel<br />

Department <strong>of</strong> Zoology,<br />

University <strong>of</strong> Florida,<br />

Gainesville,<br />

Florida 32611,<br />

U.S.A.<br />

Dr. D.K. Faulkner<br />

Entomology Department,<br />

San Diego Natural History Museum,<br />

P.O. Box 1390,<br />

San Diego,<br />

California 92112,<br />

U.S.A.<br />

Mr. G.A. Henning<br />

17 Sonderend Street,<br />

Helderkruin 1724,<br />

South Africa<br />

Mr. S.F. Henning<br />

5 Alexander Street,<br />

Florida 1709,<br />

South Africa


Dr. T. Hirowatari<br />

Entomological Laboratory,<br />

College <strong>of</strong> Agriculture,<br />

University <strong>of</strong> Osaka Prefecture,<br />

Sakai,<br />

Osaka, 591<br />

Japan<br />

Pr<strong>of</strong>. R.L. Kitching<br />

School <strong>of</strong> Australian Environmental Studies,<br />

Griffith University,<br />

Nathan,<br />

Queensland 4111,<br />

Australia<br />

Dr. J. Martin<br />

Departamento de Biologia (Zoologia),<br />

Facultad de Ciencias,<br />

Universidad Autonoma de Madrid,<br />

Madrid,<br />

Spain<br />

Dr. R.H.T. Mattoni<br />

9620 Heather Road,<br />

Beverly Hills,<br />

California 90210,<br />

U.S.A.<br />

Dr. M.C. Minno<br />

Department <strong>of</strong> Zoology,<br />

University <strong>of</strong> Florida,<br />

Gainesville,<br />

Florida 32611,<br />

U.S.A.<br />

Dr. M.L. Munguira<br />

Departamento de Biologia (Zoologia),<br />

Facultad de Ciencias,<br />

Universidad Autonoma de Madrid,<br />

Madrid,<br />

Spain<br />

Dr. D.D. Murphy<br />

Center for <strong>Conservation</strong> <strong>Biology</strong>,<br />

Stanford University,<br />

Stanford,<br />

California 94305,<br />

U.S.A.<br />

Dr. T.R. New<br />

Department <strong>of</strong> Zoology,<br />

La Trobe University,<br />

Bundoora,<br />

Victoria 3083,<br />

Australia<br />

vii<br />

Dr. M.W. Parker<br />

U.S. Fish & Wildlife Service,<br />

San Francisco Bay National Wildlife Refuge Complex,<br />

Newark,<br />

California 94560,<br />

U.S.A.<br />

Dr. J.A. Powell<br />

Department <strong>of</strong> Entomological Sciences,<br />

201 Wellman Hall,<br />

University <strong>of</strong> California,<br />

Berkeley,<br />

California 94720-0001,<br />

U.S.A.<br />

Dr. G.B. Prince<br />

Department <strong>of</strong> Lands, Parks and Wildlife,<br />

Mrs Macquarie's Road,<br />

Hobart, Tasmania 7001,<br />

Australia<br />

Dr. P.R. Samson<br />

Bureau <strong>of</strong> Sugar Experiment Stations,<br />

P.O. Box 651,<br />

Bundaberg,<br />

Qld 4650,<br />

Australia<br />

Pr<strong>of</strong>. M.J. Samways<br />

Department <strong>of</strong> Zoology and Entomology,<br />

University <strong>of</strong> Natal,<br />

Pietermaritzburg 3200,<br />

South Africa<br />

Dr. D.P.A. Sands<br />

Division <strong>of</strong> Entomology,<br />

CSIRO,<br />

Meiers Road,<br />

Indooroopilly,<br />

Qld 4068,<br />

Australia<br />

Dr. CD. Thomas<br />

School <strong>of</strong> Biological Sciences,<br />

University <strong>of</strong> Birmingham,<br />

Edgbaston,<br />

Birmingham B15 2TT,<br />

U.K.<br />

Dr. Stuart B. Weiss<br />

Center for <strong>Conservation</strong> <strong>Biology</strong>,<br />

Department <strong>of</strong> Biological Sciences,<br />

Stanford University,<br />

Stanford, California 94305,<br />

U.S.A.


PART 1. INTRODUCTION<br />

Introduction to the biology and conservation <strong>of</strong> the <strong>Lycaenidae</strong><br />

Introduction<br />

T.R. NEW<br />

Department <strong>of</strong> Zoology, La Trobe University, Bundoora, Victoria 3083, Australia<br />

The <strong>Lycaenidae</strong>, the 'blues', 'coppers', 'hairstreaks','metalmarks'<br />

and related butterflies, are the most diverse family <strong>of</strong><br />

Papilionoidea and comprise between 30 and 40% <strong>of</strong> all butterfly<br />

species. They are mostly rather small. The world's smallest<br />

butterfly may be the lycaenid Micropsyche ariana Mattoni<br />

from Afghanistan with a wingspan <strong>of</strong> only about 7mm (although<br />

some individuals <strong>of</strong> Brephidium exilis (Boisduval) can have as<br />

small a wingspan as 6mm). A few <strong>Lycaenidae</strong> are relatively<br />

large: Liphyra brassolis Westwood has a wingspan <strong>of</strong> 8–9cm,<br />

and is the largest known species. The family occurs in all<br />

major biogeographical regions in temperate and tropical zones.<br />

As with other Lepidoptera, the life cycle comprises egg,<br />

larva (caterpillar) passing through several instars, pupa and<br />

adult, with the cycle occupying several weeks to a year.<br />

Particularly in temperate regions, there may be a well-defined<br />

phenological break during winter which is passed in an inactive<br />

stage. The early stages <strong>of</strong> many taxa have been described, and<br />

a consideration <strong>of</strong> lycaenid conservation biology must include<br />

the biology <strong>of</strong> all <strong>of</strong> these life forms, from oviposition site<br />

selection by reproductive females to larval life and adult biology.<br />

In general, far more distributional and biological information is<br />

available on adults, which tend to be conspicuous and actively<br />

sought by collectors and photographers, than on the relatively<br />

inconspicuous and cryptic immature stages.<br />

Many species have very precise environmental requirements,<br />

but the family occurs in many major biomes and vegetation<br />

associations from climax forests to scrublands, grasslands,<br />

wetlands and semi-arid desert communities, many <strong>of</strong> which<br />

could be viewed as early seres in terrestrial successions. Some<br />

lycaenids have considerable potential for use as indicator species<br />

as their incidence and abundance reflects rather small degrees<br />

<strong>of</strong> habitat change.<br />

The larvae <strong>of</strong> some taxa feed on flowerbuds, flowers and<br />

fruits (Downey 1962), and thus may exert stronger selective<br />

pressures on their foodplants than many foliage feeders<br />

(Breedlove and Ehrlich 1968). Collectively, a very broad range<br />

<strong>of</strong> foods are utilised and many lycaenids have departed from the<br />

normal lepidopteran dependence on angiosperm plants to feed<br />

1<br />

on lower plants or animal material. The extent <strong>of</strong> aphytophagy,<br />

which includes predacious and mutualistic relationships with<br />

ants and various Homoptera, is sometimes both pronounced<br />

and obligatory (Cottrell 1984), so that lycaenids, as a group,<br />

participate in a wider range <strong>of</strong> ecological interactions than<br />

perhaps any other Lepidoptera.<br />

This ecological breadth has been the basis for designation <strong>of</strong><br />

'biological groups' in the family (Hinton 1951; Henning 1983).<br />

Together with the relatively comprehensive knowledge <strong>of</strong> the<br />

systematics and distribution <strong>of</strong> many temperate region taxa<br />

through longterm collector accumulation, this ecological breadth<br />

renders the family <strong>of</strong> very considerable value in conservation<br />

studies. Several species have been the targets <strong>of</strong> detailed practical<br />

measures related to their conservation in recent years, and many<br />

<strong>of</strong> these case histories are summarised in the third section <strong>of</strong> this<br />

volume.<br />

This introductory chapter enlarges on some <strong>of</strong> the topics<br />

noted above, to provide a general background to the regional<br />

and species accounts which follow.<br />

Taxonomy<br />

In this volume the <strong>Lycaenidae</strong> is taken to include the Riodininae<br />

(= Nemeobiinae) and the Styginae, both <strong>of</strong> which have been<br />

given family status by some researchers.<br />

Early classifications grouped the Riodinidae and <strong>Lycaenidae</strong><br />

s. rest, as the superfamily Lycaenoidea. Clench (1955) divided<br />

the Lycaenoidea into three families: <strong>Lycaenidae</strong> s. str., Liptenidae<br />

and Liphyridae, to which Shirozu and Yamamato (1957) added<br />

the Curetidae. In contrast, Ehrlich (1958) considered the<br />

Lycaenoidea to be a single family with the major subfamiliar<br />

groupings <strong>of</strong> Riodininae, Styginae and Lycaeninae – the latter<br />

including the four families noted in the last sentence.<br />

While the higher classification <strong>of</strong> the two 'problem' groups<br />

has proved to be controversial, the scheme proposed by Eliot<br />

(1973) (Figure la) has, with some modification, received<br />

strong support. As Eliot (1973) noted, no satisfactory<br />

classification for the whole <strong>of</strong> the <strong>Lycaenidae</strong> had been produced<br />

until then, despite notable attempts by Clench (1955) and


Figure 1. Classification <strong>of</strong> the <strong>Lycaenidae</strong> into major divisions: (a) after Eliot (1973); (b) after Ackery (1984).<br />

Stempffer (1957, 1967). Other recent authorities (such as<br />

Harvey 1987), while maintaining the Riodinidae as distinct,<br />

have allied it with the Nymphalidae. However, yet others have<br />

retained the Riodinidae and Stygidae in a somewhat broader<br />

concept <strong>of</strong> the family (Ackery 1984 and Figure lb). Thus the<br />

more recent classifications recognise eight (Eliot) or ten (Ackery)<br />

subfamilies (Figure 1). Eliot's (1973) more extensive 'tribal'<br />

divisions (Figure 2) have been more generally adopted as a<br />

working scheme by many researchers: while some <strong>of</strong> the formal<br />

divisions remain uncertain, the relationships implied appear to<br />

be valid.<br />

There is a very wide range <strong>of</strong> chromosome numbers in the<br />

family, but the modal number appears to be n = 24 (Robinson<br />

1971).<br />

Diversity and distribution<br />

The trio <strong>of</strong> Theclinae (hairstreaks), Riodininae (metalmarks)<br />

and Polyommatinae (blues) together comprise about 90% <strong>of</strong> the<br />

family. Some other subfamilies are small: Styginae, for example,<br />

(if recognised as distinct from Riodininae) includes only Styx<br />

infernalis Staudinger, from Peru. Theclinae, with well over<br />

2<br />

2000 species, is the most diverse section <strong>of</strong> the <strong>Lycaenidae</strong>.<br />

The number <strong>of</strong> species <strong>of</strong> <strong>Lycaenidae</strong> can only be estimated.<br />

In a survey Robbins (1982) produced figures <strong>of</strong> 6000–6900, an<br />

estimate which placed <strong>Lycaenidae</strong> clearly ahead <strong>of</strong> the next<br />

most diverse family, Nymphalidae. Shields (1989) noted totals<br />

<strong>of</strong> 4089 <strong>Lycaenidae</strong> s.str. and 1366 'Riodinidae' for described<br />

taxa only. Bridges (1988) listed 16,475 species-group names<br />

and more than 4000 taxonomic publications for these groups.<br />

<strong>Lycaenidae</strong> are most diverse in the tropics, especially the<br />

neotropics and southeast Asia, followed by Africa, and most<br />

species occur in tropical rainforest regions. For the neotropics,<br />

with 2650 ±350 spp., Riodininae and Theclinae are codominant,<br />

with only a few Polyommatinae, a single lycaenine, and very<br />

few others. The nearctic fauna, surprisingly, is not diverse, with<br />

only about half the number <strong>of</strong> species which occur in the<br />

Palaearctic. But, as any regional synopsis indicates, endemism<br />

at levels <strong>of</strong> both major geographical region and local community<br />

tends to be high.<br />

Very few <strong>Lycaenidae</strong> are at all widely distributed. Exceptions<br />

include Lampides boeticus (L.) which extends from Europe to<br />

Australia and Hawaii, and Celastrina argiolus (L), which Eliot<br />

and Kawazoe (1983) regard as 'one <strong>of</strong> the world's commonest<br />

and most widespread butterflies'. The latter species occurs<br />

throughout most <strong>of</strong> the Palaearctic, Oriental and Nearctic


Figure 2. Classification <strong>of</strong> the <strong>Lycaenidae</strong> (after Eliot 1973).<br />

3


egions, and has formed numerous putative subspecies over this<br />

broad range. The American subspecies have sometimes been<br />

regarded as separate species, C. ladon (Cramer) (Clench and<br />

Miller 1980).<br />

Most lycaenids are not particularly vagile and some cannot<br />

cross even small spaces between habitat patches. Although a<br />

few are occasional migrants from Europe to Britain (Lampides<br />

boeticus, Everes argiades (Pallas), Cyaniris semiargus<br />

(Rottemburg)), these are apparently rather exceptional. Small<br />

<strong>Lycaenidae</strong> have apparently crossed oceanic barriers on<br />

occasion, but this process may have played only a minor role in<br />

their evolution: Vaga Zimmerman (in the Bonins and Hawaii)<br />

and Hypojamides Riley (in Tahiti) are putative examples. L.<br />

boeticus probably reached New Zealand from Australia (Gibbs<br />

1980). A few lycaenids have been deliberately introduced to<br />

areas where they would not naturally occur. For example both<br />

Strymon bazochi Godart and Tmolus echion (L.) were introduced<br />

into Hawaii from Mexico early this century as potential agents<br />

for the control <strong>of</strong> lantana (Riotte and Uchida 1979).<br />

Poor representation <strong>of</strong> many lycaenid groups in North<br />

America led Eliot (1973) to suggest that the fauna is derived<br />

from the Old World. Tentative explanations <strong>of</strong> present<br />

distribution patterns, sometimes involving a Gondwanan origin<br />

for the group with dichotomy into Riodininae and other tribes<br />

at the time <strong>of</strong> South American separation (Eliot 1973), are not<br />

wholly convincing. Anomalies occur, perhaps related to<br />

dispersal: 'coppers' (Lycaena F.) occur in New Zealand but not<br />

in Australia or southeast Asia, and their geographically nearest<br />

congenors are in the Himalaya. Other distribution patterns are<br />

perhaps easier to explain: Udara (Vaga) blackbumi (Tuely) is<br />

one <strong>of</strong> only two endemic butterflies in Hawaii and it is thought<br />

that its ancestors may have progressively 'island-hopped' to<br />

this location. Members <strong>of</strong> another genus <strong>of</strong> small lycaenids,<br />

Zizula Chapman, may also have crossed substantial water<br />

barriers in the past; such dispersal by wind may be more<br />

frequent than commonly supposed. In Panama, Robbins and<br />

Small (1981) reported 128 species <strong>of</strong> hairstreaks being blown<br />

by the seasonal trade winds. More than 80% <strong>of</strong> these were<br />

blown through habitats where they do not normally occur.<br />

In general, though, lycaenid faunas on remote islands are<br />

small, and climatic barriers (e.g. desert) may also counter<br />

dispersal on larger land masses. Colonisation <strong>of</strong> new habitats is<br />

clearly facilitated by the presence <strong>of</strong> larval foodplants and<br />

suitable ants, but there are few quantitative data on natural<br />

colonisation <strong>of</strong> islands or other 'new' habitats. However, on the<br />

Krakatau Islands (Indonesia), 24 species have been recorded<br />

since the sterilising volcanic eruptions <strong>of</strong> 1883: none were<br />

found in 1908, seven from 1919–1921, eight by 1928–1934,<br />

and 23 were present in the 1980s (New et al. 1988). Very few<br />

<strong>of</strong> these are habitual deep forest dwellers, and perhaps 15–18 <strong>of</strong><br />

the species currently present there depend on early successional<br />

Ipomoea pes-caprae associations on accreting beach<br />

environments. Preservation <strong>of</strong> these rather transient associations<br />

on the islands appears to be a key theme to facilitate lycaenid<br />

colonisation from Java and/or Sumatra until more diverse<br />

vegetation is present. The eight species which have colonised<br />

4<br />

Anak Krakatau, last sterilised by volcanic activity in 1952, are<br />

all associated with such 'strand-line' vegetation. Natural<br />

opportunities to detect colonisation patterns <strong>of</strong> <strong>Lycaenidae</strong> on<br />

this scale are indeed rare.<br />

Life histories and biology<br />

Many species <strong>of</strong> <strong>Lycaenidae</strong> have very precise environmental<br />

needs, and a number <strong>of</strong> recent studies (especially in temperate<br />

regions) show that they may have considerable value as<br />

environmental indicators and, thus, an enhanced role in<br />

conservation studies. At this stage, though, the biology <strong>of</strong> most<br />

tropical taxa is almost entirely unknown, and a tendency to<br />

focus on the better understood temperate region forms is<br />

inevitable.<br />

Myrmecophily<br />

The life histories <strong>of</strong> slightly over 900 species (Downey 1962<br />

noted 838 species) have been documented to varying extents. A<br />

remarkable feature <strong>of</strong> many <strong>of</strong> these is dependence on ants: 245<br />

<strong>of</strong> the species noted by Downey had myrmecophilous larvae,<br />

and this dependency has attracted much attention. A tentative<br />

classification <strong>of</strong> different degrees <strong>of</strong> myrmecophily was<br />

proposed by Fiedler (1991a, 1991b).<br />

The possible advantages <strong>of</strong> myrmecophily have been<br />

addressed by, inter al., Henning (1983), Pierce (1984) and De<br />

Vries (1991 a), and a broader overview <strong>of</strong> aphytophagy is given<br />

by Cottrell (1984). In general, the ants tend to protect the<br />

caterpillars from natural enemies (Pierce and Mead 1981;<br />

Pierce and Easteal 1986) and gain additional food from caterpillar<br />

secretions (Fiedler and Maschwitz 1988). Some <strong>of</strong> these<br />

mutualisms are very complex. In Panama, the ant Ectatomma<br />

ruidum protects caterpillars <strong>of</strong> Thisbe irenea (Stoll) (Riodininae)<br />

from attacks by predatory wasps, but not from tachinid fly<br />

parasitoids, for example (De Vries 1991b).<br />

The caterpillars <strong>of</strong> T. irenea have an elaborate 'calling'<br />

system, and attract ants by producing sounds (De Vries 1990).<br />

Similar noises were recorded in other species, all <strong>of</strong> which<br />

depended on associations with ants to enhance their well being.<br />

The noises mimic vibrations used by the ants in their own<br />

communications. Stridulation is also well known in the pupae<br />

<strong>of</strong> many lycaenids, and one function <strong>of</strong> pupal sound production<br />

might also be to attract ants (Downey 1966).<br />

Females <strong>of</strong> Hypochrysops ignitus (Leach) are attracted to<br />

colonies <strong>of</strong> ant-tended Membracidae (Sands 1986). Those <strong>of</strong><br />

Allotinus major Felder & Felder lay eggs near or on membracids<br />

(probably a species <strong>of</strong> Terentius), using brooding adult female<br />

membracids as oviposition cues, and the caterpillars eat the<br />

membracid nymphs (Kitching 1987). Younger caterpillars sit<br />

on the membracids to feed and are thus exposed – unlike larvae<br />

<strong>of</strong> Feniseca Grote and Taraka Doherty which form silken<br />

retreats amongst their prey (Edwards 1886; Iwase 1953). Both<br />

the membracid and the caterpillars <strong>of</strong> A. major are tended by


workers <strong>of</strong> the same ant, Anoplolepis longipes (Jerdon) (Kitching<br />

1987).<br />

In many cases, relationships between myrmecophily or<br />

aphytophagy and oviposition patterns are not as clear as in<br />

Allotinus Felder & Felder. Laying eggs in clusters ('clustering')<br />

in many Australian <strong>Lycaenidae</strong> is strongly associated with<br />

obligate myrmecophily (Kitching 1981), but this is not the case<br />

for neotropical Riodininae, in which myrmecophilous species<br />

lay eggs singly (Callaghan 1986). In the latter group, clustering<br />

is associated with gregarious behaviour. Myrmecophilous<br />

riodinine larvae are solitary, and gregarious larvae are not<br />

myrmecophilous. The latter may be aposematic and conspicuous,<br />

so that their protection against many predators is a function <strong>of</strong><br />

their distastefulness. Many myrmecophilous larvae, in contrast,<br />

are rather cryptic, and accompanying ants may help to prevent<br />

them being attacked by predators and parasitoids.<br />

Such relationships between caterpillars and ants are thus <strong>of</strong><br />

central importance in considering the evolution and biology <strong>of</strong><br />

<strong>Lycaenidae</strong> and have attracted much attention.<br />

Myrmecophily and evolution in lycaenids<br />

Symbiosis with ants may have been an early development in the<br />

evolution <strong>of</strong> <strong>Lycaenidae</strong> (Eliot 1973), and both Hinton (1951)<br />

and Malicky (1969) suggested that ancestral lycaenids were<br />

myrmecophilous. Contrary to Pierce's (1987) conclusion that<br />

the distribution <strong>of</strong> myrmecophily in <strong>Lycaenidae</strong> does not reflect<br />

phylogeny, Fiedler (1991a, 1991b) believed that there was a<br />

strongly phylogenetic relationship present.<br />

However, there is some possible confusion over roles <strong>of</strong><br />

myrmecophily in lycaenoid evolution as their influences on<br />

Riodinines and the other taxa may be markedly different (De<br />

Vries 1991a). Not only are they the most common basis for<br />

suggesting ecological groupings in the family (Henning 1983),<br />

but the evolution <strong>of</strong> lycaenid diversity itself may also be<br />

involved. Pierce (1984) suggested that lycaenid diversity may<br />

reflect speciation in relation to other butterfly families, and that<br />

this could be influenced by larvae/ant associations in two<br />

important ways:<br />

1. Female lycaenids may adopt ants as oviposition cues<br />

(Fiedler and Maschwitz 1989, on Anthene emolus (Godart)) so<br />

that the presence <strong>of</strong> ants on a novel foodplant may induce a<br />

rapid host switch. Although few such 'oviposition mistakes'<br />

(Pierce 1984) may actually lead to range extensions, it may be<br />

more important for a given lycaenid to retain a particular ant<br />

association than a particular foodplant, and an increase in the<br />

number <strong>of</strong> ovipositions on different foodplants may increase<br />

the number <strong>of</strong> opportunities for subsequent speciation.<br />

Essentially, novel foodplant choices may be made by female<br />

lycaenids to an unusually high degree because they select for<br />

ants as well as for chemically and physically suitable foodplants.<br />

A 'new' hostplant may occupy a different ecological range<br />

from those utilised earlier, and population isolates could thus be<br />

formed.<br />

2. The general non-vagility <strong>of</strong> many lycaenids results in<br />

their occurrence in small, semi-isolated populations with rather<br />

5<br />

little regular genetic interchange between them. Pierce (1983)<br />

showed that a deme <strong>of</strong> the Australian Jalmenus evagoras<br />

(Donovan) may be restricted to a single Acacia tree, where<br />

males aggregate and compete for emerging females so that<br />

variability in male reproductive success effectively reduces<br />

population size further. Such patchy distributions (also noted in<br />

the North American Glaucopsyche lygdamus Doubleday: Pierce<br />

1984) occur in spite <strong>of</strong> apparent continuous foodplant availability<br />

and it is quite possible that they result from selection <strong>of</strong><br />

foodplant areas which are high in nitrogen, as well as having the<br />

required ants. Many myrmecophilous lycaenid larvae actively<br />

prefer nitrogen-rich foodplants and plant parts such as seed<br />

pods and flowers. This may be explained in part by the need to<br />

supply ants with amino acids as a 'nutrient reward' for tending<br />

the larvae (Pierce 1984).<br />

Larval feeding<br />

The overall importance <strong>of</strong> plant-feeding to caterpillars <strong>of</strong><br />

<strong>Lycaenidae</strong> differs substantially between different subfamilies,<br />

and those <strong>of</strong> some groups rarely take plant food. As far as is<br />

known, all species <strong>of</strong> Poritiinae and Lycaeninae are normally<br />

phytophagous. Some Curetinae are phytophagous. Lipteninae<br />

are also plant feeders, but are highly unusual amongst butterflies<br />

in that larval food usually consists <strong>of</strong> algae, fungi or lichens (see<br />

Cottrell 1984, for summary). Most genera <strong>of</strong> the two largest<br />

subfamilies, Theclinae and Polyommatinae, appear to be<br />

phytophagous or opportunistically carnivorous with varying<br />

degrees <strong>of</strong> dependence on prey. Maculinea van Eecke and<br />

Lepidochrysops Hedicke larvae are phytophagous when young,<br />

but the late instars are obligate predators <strong>of</strong> ant larvae. Other<br />

aphytophagous genera are noted in Table 1. Both Liphyrinae<br />

and Miletinae appear to be entirely aphytophagous, and the<br />

unlisted genera in Table 1 reflect ignorance <strong>of</strong> their larval<br />

biology, rather than known phytophagy. Liphyra Westwood<br />

and Euliphyra Holland larvae are probably specific feeders on<br />

early stages <strong>of</strong> tree ants (Oecophylla spp): their larvae are<br />

flattened and have a heavily armoured cuticle which enables<br />

them to withstand ant attacks. The pupa <strong>of</strong> Liphyra remains<br />

inside the last larval skin, which thereby functions as a puparium.<br />

Aslauga larvae are predators <strong>of</strong> Homoptera, at least as late<br />

instars. Eggs <strong>of</strong> Miletinae are typically laid near colonies <strong>of</strong><br />

Homoptera, including aphids, coccids and membracids and<br />

some, at least, are found on a wide range <strong>of</strong> different hostplants.<br />

Although the larvae are predominantly predators, some younger<br />

instars may also feed on honeydew or other insect secretions<br />

such as aphid cornicle secretions.<br />

Selection <strong>of</strong> foodplant species by phytophagous species,<br />

and their effects on foodplants, are difficult to study. Flower<br />

predation <strong>of</strong> a range <strong>of</strong> perennial herbaceous legumes by<br />

Glaucopsyche lygdamus in Colorado differed substantially<br />

between species (Breedlove and Ehrlich 1972), with either<br />

Lupinus or Theropsis being by far the most heavily attacked<br />

plant at each <strong>of</strong> a series <strong>of</strong> sites. On both plant genera, flowerfeeding<br />

can markedly reduce seed-set (Breedlove and Ehrlich<br />

1968,1972). Whereas G. lygdamus females select inflorescences


Table 1. <strong>Lycaenidae</strong> with larvae which do not feed on plants (after Cottrell 1984, based on Eliot 1973).<br />

Subfamily<br />

Lipteninae<br />

Poritiinae<br />

Liphyrinae<br />

Milelinae<br />

Curetinae<br />

Theclinae<br />

Lycaeninae<br />

Polyommatinae<br />

Number <strong>of</strong><br />

Tribes Genera<br />

2<br />

1<br />

1<br />

4<br />

1<br />

19<br />

1<br />

4<br />

46<br />

7<br />

5<br />

12<br />

1<br />

241<br />

18<br />

149<br />

Aphytophagous genera*<br />

(none)<br />

(none)<br />

(3) Liphyra, Euliphyra, Aslauga<br />

(8) Miletus, Allotinus, Megalopalpus, Taraka, Spalgis, Feniscea,<br />

Lachnocnema, Theslor<br />

(none)<br />

(7) Acrodipsas, Shirozua, Zesius, Spindasis, Oxychaela, Trimenia,<br />

Argyrocupha<br />

(none)<br />

(4) Triclena, Niphanda, Maculinea, Lepidochrysops<br />

• Members <strong>of</strong> some genera are partially phytophagous, some only in the early instars (e.g. Maculinea, Lepidochrysops).<br />

<strong>of</strong> less hairy Lupinus species on which to lay, the reverse trend<br />

is true <strong>of</strong> Plebejus icarioides Boisduval, which oviposits on the<br />

most hairy (non-floral parts) <strong>of</strong> Lupinus by preference (Downey<br />

1962). Some caution is needed in extrapolating from laboratory<br />

feeding trials to the field - thus, captive larvae <strong>of</strong> P. icarioides<br />

will feed on any species <strong>of</strong> Lupinus pr<strong>of</strong>fered (Downey and<br />

Fuller 1961) but wild populations normally utilise only a few <strong>of</strong><br />

the species available in their habitat. There may also be a<br />

pronounced seasonal variation in foodplant quality: P. acmon<br />

(Westwood) and related species utilise some foodplants only at<br />

particular times <strong>of</strong> the year (Goodpasture 1974). J. evagoras<br />

females preferred to oviposit on potted Acacia which had been<br />

given nitrogenous fertiliser rather than on similar plants given<br />

water alone (Baylis and Pierce 1991).<br />

Comparatively few species <strong>of</strong> <strong>Lycaenidae</strong> seem to be<br />

markedly polyphagous and many clearly have very restricted<br />

ranges <strong>of</strong> foodplants. Broad taxonomic ranges <strong>of</strong> foodplants –<br />

such as the 30 or so species (representing the families<br />

Selaginaceae, Lamiaceae, Verbenaceae, Fabaceae and<br />

Geraniaceae) recorded for Lepidochrysops in Africa (Cottrell<br />

1984) – are commonly recorded only for taxa which are later<br />

ant-feeders. However, there is evidence <strong>of</strong> more local host<br />

restriction for Lepidochrysops, so that in any one area only a<br />

single plant genus is used for oviposition (Cottrell 1984). In<br />

general support <strong>of</strong> this relationship between poly phagy and ant<br />

attendance, a broad foodplant range for Hypochrysops miskini<br />

(Waterhouse) in Queensland led Valentine and Johnson (1989)<br />

to suggest that rainforest lycaenids may be polyphagous because<br />

a narrow choice <strong>of</strong> plants may be restrictive and cancel out<br />

advantages <strong>of</strong> ant-attendance. In Queensland, rainforest species<br />

confined to single host plant species tend not to be ant-attended.<br />

Rarely, different generations <strong>of</strong> the same species may utilise<br />

different foodplants: the first (spring) generation <strong>of</strong> Celastrina<br />

argiolus (L.) in Europe feeds on holly, while the second<br />

(summer) generation larvae eat ivy. This species also shows<br />

6<br />

marked seasonal dimorphism, but it is not clear if this is foodrelated.<br />

Nitrogen-rich plant feeding is a characteristic <strong>of</strong> many<br />

lycaenid taxa. Pierce (1985) has noted that caterpillars <strong>of</strong><br />

lycaenids from Australia, South Africa and North America<br />

have been recorded feeding on most known non-leguminous,<br />

nitrogen-fixing plant families. Several species <strong>of</strong> Cycadaceae<br />

are included, and most Lipteninae are probably specialised<br />

lichen-feeders. There are, <strong>of</strong> course, exceptions: the nonmyrmecophilous<br />

riodinine Sarota gyas (Cramer) in Panama<br />

and Costa Rica feeds exclusively on epiphyUs (De Vries 1988),<br />

and those epiphylls tested were not nitrogen-fixers. Females <strong>of</strong><br />

this species lay eggs on taxonomically disparate hostplants with<br />

old leaves covered by epiphylls. In many other lycaenids,<br />

though, ovipositions on reproductive structures <strong>of</strong> larval<br />

foodplants may reflect selection for high nitrogen levels in the<br />

future larval food.<br />

Feeding on floral structures may itself engender polyphagy:<br />

Robbins and Aiello (1982) noted that the thecline Strymon<br />

melinus (Hübner) is one <strong>of</strong> the most polyphagous butterfly<br />

species known, and has been recorded feeding on flowers <strong>of</strong> 46<br />

genera in 21 families <strong>of</strong> plants. Members <strong>of</strong> the genus Callophrys<br />

Billberg also feed on plants <strong>of</strong> some 11 families (Pratt and<br />

Ballmer 1991). Polyphagy <strong>of</strong> this sort might be an important<br />

ecological strategy for the lycaenids in exploiting and adjusting<br />

to changing environmental conditions and could be a correlate<br />

<strong>of</strong> r-selection. Flower-feeding could, perhaps, allow larval<br />

access to plants when the foliage is not available due to their<br />

chemical defences (Robbins and Aiello 1982), and it is likely<br />

that some lycaenids in the tropics have life cycles which are<br />

well adapted to exploit peak flowering seasons <strong>of</strong> putative<br />

foodplants. Robbins and Aiello cite data (Croat 1978) that peak<br />

flowering on Barro Colorado Island (Panama) occurs at the end<br />

<strong>of</strong> the dry season, a period corresponding with a marked<br />

increase in abundance <strong>of</strong> <strong>Lycaenidae</strong> but a decrease <strong>of</strong> many


other butterflies. Such seasonal apparency may be an important<br />

feature facilitating the use <strong>of</strong> tropical lycaenids as indicators <strong>of</strong><br />

habitat quality. It may also facilitate ecologically important<br />

dispersal, as the butterflies are abundant at the time <strong>of</strong> strong<br />

sustained dry season trade winds (Robbins and Small 1981).<br />

Samples were skewed heavily toward females, and many <strong>of</strong> the<br />

128 species were being blown through habitats where they do<br />

not occur normally.<br />

Most phytophagous lycaenid larvae are external feeders,<br />

either eating whole plant tissues or organs, or skeletonising<br />

foliage. However, Callophyrysxami (Reakirt) in Mexico feeds<br />

on fleshy-leaved Crassulaceae and, on Echeveria gibbiflora,<br />

caterpillars may burrow completely into the leaves and feed on<br />

the fleshy internal pulp. Wastes are expelled through the entrance<br />

hole (Ziegler and Escalante 1964). Larvae <strong>of</strong> some other<br />

Callophrys (such as C. viridis (Edwards)) feed on flowers <strong>of</strong><br />

Eriopsis and, as with other larvae with similar habits, they<br />

resemble their foodplant flowers in coloration (Brown and<br />

Opler 1967). Variation in larval coloration occurs in other taxa,<br />

such as Zizina labradus (Godart) in Australia (Sibatani 1984).<br />

Although it is clear that very considerable trophic specialisation<br />

occurs in many lycaenids – both in phytophagous and<br />

aphytophagous species where intricate and obligate relationships<br />

with other animals are common – little is known <strong>of</strong> the biology<br />

<strong>of</strong> most species <strong>of</strong> the family. The precise resource needs <strong>of</strong> many<br />

specialised species cannot yet be appreciated fully, but they are<br />

likely to render the species very susceptible to habitat change and<br />

consequent changes in resource supply.<br />

Adult feeding<br />

Adult lycaenids (<strong>of</strong> Theclinae, Lycaeninae and Polyommatinae)<br />

commonly seek nectar from flowers, although some Theclinae<br />

may depend more on aphid honeydew. In contrast, Lipteninae<br />

and Miletinae are not known to visit flowers, and depend on<br />

secretions from extrafloral nectaries or on homopteran<br />

honeydew. Adult Liphyra lack a proboscis and thus cannot<br />

feed. A few taxa may feed on secretions from the dorsal nectary<br />

organ <strong>of</strong> other ant-tended lycaenid larvae.<br />

Adult sexual dimorphism<br />

This is widespread in the family. Commonly, males are brighter<br />

(either the females are browner or have wider, dark borders to<br />

the wings). Indeed, in some taxa it is difficult to associate the<br />

sexes correctly because <strong>of</strong> relatively extreme dimorphism.<br />

Males may have modified scales or sex-brands on either surface<br />

<strong>of</strong> the forewing or on the upperside <strong>of</strong> the hindwing. These may<br />

be associated with hair-tufts but, more commonly, distinct sex<br />

brands are absent and androconia are scattered irregularly over<br />

the upper surface <strong>of</strong> the wings.<br />

Adult behaviour<br />

Males <strong>of</strong> some species are regular hill-toppers: some species <strong>of</strong><br />

the 'Lycaenopsis-group', for example, are rarely seen elsewhere,<br />

7<br />

and both sexes <strong>of</strong> some species <strong>of</strong> Monodontides Toxopeus<br />

appear to be hill-top residents (Eliot and Kawazoe 1983). Hilltopping<br />

may be common in riodinines (Shields 1967).<br />

Protandry may occur regularly in some lycaenids. Males <strong>of</strong><br />

Polyommatus icarus (Rottemburg) not only emerge before the<br />

females but also fly more strongly. Males <strong>of</strong> this genus interact<br />

strongly with other males, and it seems that the bright male<br />

colours may indeed be 'directed' at other males. Lundgren<br />

(1977) made the intriguing suggestion that the high frequency<br />

<strong>of</strong> the inter-male encounters may aid in regulating dispersal and<br />

population density, and this may occur also between groups <strong>of</strong><br />

sympatric species with similar appearance.<br />

Displays and pheromones both appear to be involved in<br />

ensuring conspecific matings in lycaenids living in the same<br />

area. The female anal hair-tuft <strong>of</strong> Nordmannia Tutt may play a<br />

role in courtship (Nakamura 1976). Multiple copulations can<br />

occur (Fujii, 1989, on Shirozua janasi (Janson)), and it is<br />

possible that copulatory mate guarding occurs in this thecline.<br />

The time <strong>of</strong> day for mating differs between species: S. janasi<br />

copulates during the day whereas two species <strong>of</strong> Japonica Tutt<br />

mate at dusk (Fujii 1989). There are occasional records <strong>of</strong><br />

cross-taxon matings, sometimes even between subfamilies<br />

(Shapiro 1985), or even families (Shapiro 1982; Johnson 1984),<br />

usually between phenotypically similar butterflies.<br />

'Perching' behaviour appears to be an important isolating<br />

mechanism in Riodininae (Callaghan 1979,1983). Neotropical<br />

riodinines appear to have developed a range <strong>of</strong> different perching<br />

strategies which help to maintain specific isolation within the<br />

habitat. Habitat complexity is an important facet <strong>of</strong> this aspect<br />

<strong>of</strong> these forest butterflies, as the spacing <strong>of</strong> perching sites and<br />

positions <strong>of</strong> perching differ considerably between genera – by<br />

topographic features, by light/shadow regimes and by time <strong>of</strong><br />

day. Following earlier accounts by Scott (1968) and Shields<br />

(1967), Callaghan suggested that perching (and hill-topping)<br />

species tend to have low population densities, and 'rendezvous'<br />

localities help such rarer species to find mates. Perching for<br />

short periods (an average <strong>of</strong> 2.4h in 10 riodinine genera)<br />

minimises exposure to predators. Different perching strategies<br />

are enhanced by ethological strategies such as displays. When<br />

males are scarce females may actively search out perching sites<br />

and await them if they are ready to mate. Ethological isolating<br />

mechanisms <strong>of</strong> this sort may be more important in taxa which<br />

are not strict deep forest dwellers.<br />

Perching behaviour <strong>of</strong> different species <strong>of</strong> Cupressaceaefeeding<br />

Callophrys is distinctly non-random (Johnson and<br />

Borgo 1976). The insects apparently orientate to the position <strong>of</strong><br />

the sun, and taller trees are most <strong>of</strong>ten selected for perching on.<br />

Perching posture in lycaenids, as in other butterflies, may have<br />

an important role in thermoregulation (Clench 1966).<br />

<strong>Lycaenidae</strong> which are normally 'perchers' rather than<br />

'patrollers' may at times undergo sustained flights and the<br />

down-valley flights <strong>of</strong> four such species <strong>of</strong> Theclini in Colorado<br />

and Arizona (Scott 1973) appeared to be a food-seeking<br />

behaviour. This may also be so for Satyrium Scudder in California<br />

(MacNeill 1967). Some species fly only at particular times <strong>of</strong><br />

the day and, within a genus, species co-occurring at the same


locality may not overlap in their flight periodicity (see Sibatani<br />

1992, on Favonius Sibatani & Ito).<br />

Some very small blues are extremely weak fliers, as suggested<br />

earlier. Some have elongated narrow wings which, at least in<br />

Zizula hylax (F.), enable the butterflies to crawl into the corollas<br />

<strong>of</strong> long-tubed flowers and obtain nectar (Ehrlich and Ehrlich<br />

1972).<br />

Defence against predators<br />

It has been suggested that a few adult <strong>Lycaenidae</strong> participate in<br />

mimicry complexes, either with other blues or other kinds <strong>of</strong><br />

Lepidoptera. I have netted specimens <strong>of</strong> Udara Toxopeus in<br />

stream beds in New Guinea in the belief that they were small<br />

members <strong>of</strong> one <strong>of</strong> the many species <strong>of</strong> Delias Hübner (Pieridae)<br />

in the same habitat; Plebejus icarioides may be a model for the<br />

noctuid moth Caenurgina caerulea Grote in North America<br />

(Downey 1965); and male Lycaena heteronea Boisduval may<br />

be a mimic <strong>of</strong> G. lygdamus and other sympatric blues (Austin<br />

1972). The latter feed on plants containing alkaloids (Lupinus)<br />

or selenium (Astragalus), while the foodplants <strong>of</strong> the Lycaena<br />

(Eriogonum) apparently do not. However, experimental<br />

investigations <strong>of</strong> this scenario do not appear to have been made.<br />

In Sierra Leone, Pseudaletis leonis (Staudinger) may be a<br />

mimic <strong>of</strong> a similarly-patterned diurnal arctiid moth, or <strong>of</strong> a<br />

larger danaine (Owen 1991).<br />

Other putative defences against predators include the<br />

chemical defence <strong>of</strong> Eumaeus atala (Poey). This species<br />

sequesters a deterrent defensive chemical (cycasin) from larval<br />

cycad foodplants, and this clearly renders the adults unpalatable<br />

to birds (Bowers and Larin 1989). Cycasin also deters attack by<br />

Camponotus ants. Adult E. atala are aposematic, and the genus<br />

is mimicked by several other Lepidoptera (including species <strong>of</strong><br />

Castniidae and Noctuidae) in South America. Further examples<br />

<strong>of</strong> defence can be found at all growth stages: the 'false heads',<br />

including hindwing tails <strong>of</strong> many adults (Robbins 1980,1981);<br />

the 'monkey head' pupae <strong>of</strong> Spalgis Moore (Hinton 1974);<br />

sound production by pupae <strong>of</strong> some taxa; unusual hairiness or<br />

thickened cuticle <strong>of</strong> some caterpillars; and the female anal hairtuft<br />

<strong>of</strong> some species <strong>of</strong> Nordmannia which is used to cover eggs<br />

with long scales at the time <strong>of</strong> oviposition.<br />

Even mode <strong>of</strong> hatching from the egg may be correlated with<br />

protection from natural enemy attack: caterpillars <strong>of</strong> Maculinea<br />

alcon (Denis & Schiffermüller) and M. rebeli (Hirschke) hatch<br />

through the base <strong>of</strong> the eggshell and emerge on the opposite side<br />

<strong>of</strong> the leaf (Thomas et al. 1991). The 'normal' exposed eggshells<br />

are unusually thick, perhaps to counter enemy attack.<br />

Particularly in the tropics, lycaenid biology is almost<br />

unstudied at other than the most superficial levels. Most<br />

information, including that which has led to tentative<br />

generalisations for the family, has been derived from temperate<br />

region species. Even for the best-studied lycaenid faunas, there<br />

are many gaps in fundamental biological knowledge.<br />

8<br />

Endangering processes<br />

In their review <strong>of</strong> threatened swallowtail butterflies, Collins<br />

and Morris (1985) enumerate a number <strong>of</strong> processes which may<br />

lead to the decline <strong>of</strong> these butterflies, and it is instructive to<br />

assess the effects <strong>of</strong> these processes on the ecologically very<br />

different <strong>Lycaenidae</strong>.<br />

Collecting and trade<br />

The effects <strong>of</strong> collecting are controversial and difficult to<br />

evaluate fully. Many lycaenids occur in small closed populations<br />

and because <strong>of</strong> this may be much more vulnerable to localised<br />

collector pressure than many other butterflies. There is at least<br />

some suggestion, for example, that the demise <strong>of</strong> the Large<br />

Copper, Lycaena dispar (Haworth), in Britain in the midnineteenth<br />

century was in part due to increased collector pressure<br />

on populations which had been rendered vulnerable by habitat<br />

destruction. In most cases collecting is probably the subsidiary<br />

rather than the prime cause <strong>of</strong> decline or extinction. Perhaps,<br />

especially in northern temperate regions, there is reason to<br />

suggest that this syndrome could occur for many localised taxa<br />

sought by collectors who visit the same colony year by year.<br />

Even specialist collecting to help monitoring <strong>of</strong> restricted<br />

colonies for conservation assessment may cause direct damage<br />

to populations: Murphy (1989) has recently pointed out that<br />

conventional mark-recapture studies used to estimate population<br />

sizes <strong>of</strong> taxa <strong>of</strong> conservation interest may cause inadvertent<br />

damage, either by mutilation through handling, or by inducing<br />

changes in individual behaviour. Small delicate butterflies,<br />

such as most lycaenids, may be particularly vulnerable to such<br />

effects.<br />

Commercial collecting is <strong>of</strong> considerably less importance<br />

for <strong>Lycaenidae</strong> than for large showy butterflies. Rare species<br />

from unusual localities will always find a market amongst<br />

wealthy collectors and museums, but the 'ornament' and<br />

'souvenir' trades in <strong>Lycaenidae</strong> are very low. In Malaysia, for<br />

example, the vast bulk <strong>of</strong> this trade is in Papilionidae and<br />

Nymphalidae: in stores in Kuala Lumpur in 1988,I noticed only<br />

very few <strong>Lycaenidae</strong> (all relatively large species <strong>of</strong> Arhopala<br />

Boisduval and related genera). In general, for this trade, 'small'<br />

is distinctly not 'beautiful' or desirable.<br />

Similarly, <strong>Lycaenidae</strong> are not particularly desired as exhibits<br />

in butterfly houses, and the difficulty <strong>of</strong> rearing many <strong>of</strong> the<br />

species may also be a deterrent in this context. Collins (1987a)<br />

lists only eight species <strong>of</strong> <strong>Lycaenidae</strong> (<strong>of</strong> 223 butterfly species<br />

in total) on show in a sample <strong>of</strong> 18 butterfly houses in Britain<br />

in 1986 (Table 2). Seven <strong>of</strong> these were obtained as pupae, but<br />

the Malaysian Spindasis syama (Horsfield) were apparently<br />

obtained as adults by the one exhibitor.<br />

A total collecting ban is difficult to enforce, especially on<br />

unreserved public land – indeed, it is impossible without<br />

wardenship or other constant (and probably expensive) security<br />

measures. Even responsible, private collectors obeying voluntary<br />

restrictive quota codes may cause harm if they are in sufficient


numbers. As Collins and Morris (1985) commented for<br />

Papilionidae 'there is a danger that collectors may be unable to<br />

recognise when they are depleting butterfly stocks below the<br />

threshold <strong>of</strong> recovery, particularly when they only visit the<br />

breeding area for short periods <strong>of</strong> time'. Although no <strong>Lycaenidae</strong><br />

are currently listed in CITES Appendices, a number <strong>of</strong> rare or<br />

local species have received local legislative protection (if not<br />

more tangible conservation) through bans on collecting. In<br />

addition, over 40 species are mentioned in various countrybased<br />

European legislation (Collins 1987b and Table 3) and 28<br />

taxa are listed in the United States Federal Register <strong>of</strong> Endangered<br />

and Threatened Wildlife. The most prolific, and probably the<br />

least discriminating legislation is in India, where some 160<br />

species are listed under the Wildlife Protection Act (Table 4).<br />

The Code <strong>of</strong> <strong>Conservation</strong> Responsibility adopted by<br />

commercial entomologists in Britain in 1974, restricts trade in<br />

a number <strong>of</strong> species (including Maculinea arion (L.), then still<br />

extant in Britain) to specimens already in 'circulation', so that<br />

the only legal way that a collector can purchase examples <strong>of</strong><br />

these species is from the limited pool already in collections.<br />

Seventy-nine lycaenid taxa arc included in the <strong>IUCN</strong> Red<br />

List <strong>of</strong> Threatened Animals (1990) (Table 5). Many <strong>of</strong> these are<br />

also cited in various country-based legislations.<br />

Habitat alteration and destruction<br />

This is the prime threat to all lycaenid species with limited<br />

distributions and low vagility, and has already been the agent <strong>of</strong><br />

major declines <strong>of</strong> many <strong>of</strong> these – as the examples discussed<br />

later in this volume attest. Lycaenid taxa are particularly<br />

susceptible to certain types <strong>of</strong> habitat alteration including:<br />

changes in forestry practices in tropical and temperate regions;<br />

conversion <strong>of</strong> shrublands to pasture and agricultural lands;<br />

wetland drainage; heathland succession; grassland management<br />

practices; the effects <strong>of</strong> grazing animals such as rabbits; and<br />

expanding urban, industrial and recreational land use.<br />

In common with all other animals and plants, levels <strong>of</strong><br />

concern therefore range from large-scale destruction <strong>of</strong> tropical<br />

rainforests whose biota have scarcely been documented (and in<br />

many cases never now can be), to small, local habitat changes<br />

Table 2. <strong>Lycaenidae</strong> flown in butterfly houses in Britain (Collins 1987a).<br />

Species<br />

Eumaeus alula<br />

I ampules boeticus<br />

Lycaena helle<br />

L. phlaeas<br />

Narathura centaurus<br />

Polyommatus icarus<br />

Spindasis syama<br />

Telicada nyseus<br />

Native range<br />

Caribbean<br />

(widespread)<br />

Europe, Asia<br />

Europe, Asia<br />

Malaysia<br />

Europe<br />

Asia<br />

East Asia<br />

9<br />

in 'ordinary' vegetation such as grassland or heathland. Such<br />

changes have, <strong>of</strong> course, occurred in many parts <strong>of</strong> the world,<br />

and in many major areas their effects have been both unheralded<br />

and undocumented so that we can merely infer, from the present<br />

status <strong>of</strong> taxa and knowledge <strong>of</strong> their biology as this accumulates,<br />

the magnitude <strong>of</strong> their effects. Only rarely outside northern<br />

temperate regions has conservation awareness for <strong>Lycaenidae</strong><br />

progressed to the stage where concern can be shown in any<br />

practical manner. There, it is sometimes abundantly clear that<br />

decline <strong>of</strong> species or assemblages, and the initiation <strong>of</strong> concern<br />

for their conservation, has been engendered by particular<br />

localised human activities – sometimes as 'one-<strong>of</strong>f destructive<br />

events, sometimes more broadly. Many <strong>of</strong> the former apply to<br />

remnant populations which represent formerly much more<br />

widespread taxa which have become progressively vulnerable<br />

over a longer time. In other parts <strong>of</strong> the world, very many<br />

species have been recorded only from single or highly disjunct<br />

localities, and even sound demonstration <strong>of</strong> their abundance or<br />

dependence on particular habitats is difficult or impossible at<br />

this time. In short, status evaluation is difficult or impossible,<br />

and the option <strong>of</strong> habitat protection, in the interest <strong>of</strong> decreasing<br />

or eliminating perceived threats, is the most urgent option.<br />

In contrast to most Papilionidae, grassland and other open<br />

vegetation types are vitally important habitats to many<br />

<strong>Lycaenidae</strong>. In Europe, calcareous grassland is a particularly<br />

important lycaenid habitat which has undergone large scale and<br />

sometimes dramatic changes. The extinction <strong>of</strong> Cyaniris<br />

semiargus (Rottemburg) in Britain as long ago as 1877 has been<br />

attributed to changes in grassland management (Heath 1981).<br />

While 'traditional' methods <strong>of</strong> land and vegetation<br />

maintenance, such as coppicing <strong>of</strong> forests, may foster the wellbeing<br />

<strong>of</strong> some species, intensification <strong>of</strong> agricultural practices<br />

has caused concern for some. Wetlands are particularly<br />

vulnerable to such changes and a number <strong>of</strong> European wetland<br />

<strong>Lycaenidae</strong> are under threat. Some species <strong>of</strong> Maculinea and<br />

Lycaena restricted to this habitat are particularly endangered.<br />

Drainage <strong>of</strong> the fens in England last century was a prime cause<br />

<strong>of</strong> decline <strong>of</strong> Lycaena dispar.<br />

Urbanisation has caused concern for lycaenids in places as<br />

far apart as Los Angeles, California and Melbourne, Australia.<br />

Origin <strong>of</strong> material<br />

U.S.A.: Florida<br />

Sri Lanka, Malaysia<br />

France<br />

France<br />

Malaysia<br />

U.K<br />

Malaysia<br />

Sri Lanka, Malaysia


Table 3. <strong>Lycaenidae</strong> protected by legislation in Europe (from Collins 1987b).<br />

(i) Areas which protect <strong>Lycaenidae</strong> as part <strong>of</strong> 'all butterflies' or a similar ordinance, sometimes with exceptions for particular Pieridae stated. Date <strong>of</strong> legislation<br />

given in parentheses:<br />

Austria: Niederosterreich (1978), Salzburg (1980), Steiermark (1936), Tyrol (1975), Oberosterreich (1982), Vienna (1985), Vorarlberg (1979).<br />

Germany: (Law <strong>of</strong> German Democratic Republic 1984*).<br />

Luxembourg: (1986).<br />

(ii) Particular taxa<br />

Taxon<br />

Agrodiaetus admelus<br />

A. damon<br />

A, riparlii<br />

Aricia agestis<br />

A. crassipuncta<br />

A. taberdiana<br />

Callophrys mystaphia<br />

C. suaveola<br />

Cupido osiris<br />

Cyaniris Helena<br />

Eumedonia damon<br />

Freyeria trochylus<br />

lolana iotas<br />

Kretania eurypilus<br />

K. psylorita<br />

Lycaeides argyrognomon<br />

L. idas<br />

L. helle<br />

Lysandra bellargus<br />

L. caucasica<br />

Maculinea alcon<br />

M. arion<br />

M. nausithous<br />

M. rebeli<br />

Country<br />

Greece<br />

Hungary<br />

Greece<br />

Hungary<br />

Germany (FR)<br />

Germany (FR)<br />

Germany (FR)<br />

Germany (FR)<br />

Hungary<br />

Greece<br />

Hungary<br />

Greece<br />

Hungary<br />

Germany (FR)<br />

Germany (FR)<br />

Greece<br />

Belgium: Wallone Region<br />

Belgium: Flemish Region<br />

Belgium: Wallone Region<br />

Finland<br />

France (females only)<br />

Germany (FR)<br />

Hungary<br />

Netherlands<br />

Belgium: Wallone Region<br />

France (females only)<br />

Germany (FR)<br />

Hungary<br />

Belgium: Wallone Region<br />

France (L. b. coelestis, females only)<br />

Germany (FR)<br />

Belgium: Flemish Region<br />

Germany (FR)<br />

Belgium: Flemish Region<br />

Belgium: Wallone Region<br />

Germany (FR)<br />

United Kingdom<br />

Germany (FR)<br />

Hungary<br />

Belgium: Wallone Region<br />

Germany (FR)<br />

10<br />

Date <strong>of</strong> legislation<br />

(1980)<br />

(1982)<br />

(1980)<br />

(1982)<br />

(1986)*<br />

(1986)<br />

(1986)<br />

(1986)<br />

(1982)<br />

(1980)<br />

(1982)<br />

(1980)<br />

(1982)<br />

(1986)<br />

(1986)<br />

(1980)<br />

(1985)<br />

(1980)<br />

(1985)<br />

(1976)<br />

(1979)<br />

(1986)<br />

(1982)<br />

(1973)<br />

(1985)<br />

(1979)<br />

(1986)<br />

(1982)<br />

(1985)<br />

(1979)<br />

(1986)<br />

(1980)<br />

(1986)<br />

(1980)<br />

(1985)<br />

(1986)<br />

(1981)<br />

(1986)<br />

(1982)<br />

(1985)<br />

(1986)


Table 3 (cont.). <strong>Lycaenidae</strong> protected by legislation in Europe (from Collins 1987b).<br />

Taxon<br />

M. teleius<br />

Neolycaena coelestina<br />

Nordmannia acaciae<br />

N. armenia<br />

N. marcidus<br />

N. sassanides<br />

Paleochrysophanus hippothoe<br />

Plebejus pylaon<br />

Polyommatus eros<br />

Pseudophilotes bavius<br />

Strymonidia pruni<br />

S. myrrhina<br />

Thersamonia thetis<br />

Tomares callimachus<br />

T. romanovi<br />

Turanana panagea<br />

Vacciniina optilate<br />

Zizeeria knysna<br />

Country<br />

Belgium: Flemish Region<br />

France ( M. t. burdigalensis, females only)<br />

Germany (FR)<br />

Belgium: Wallone Region<br />

Germany (FR)<br />

Germany (FR)<br />

Germany (FR)<br />

Hungary<br />

Hungary<br />

Greece<br />

Germany (FR)<br />

Greece<br />

Germany (FR)<br />

Greece<br />

Germany (FR)<br />

Germany (FR)<br />

Greece<br />

Germany (FR)<br />

Greece ( Z. k . cassandra)<br />

• FR = Federal Republic: it is not yet clear how the former laws for FRG and DDR will operate for a unified Germany.<br />

Table 4. <strong>Lycaenidae</strong> listed in The Indian Wildlife (Protection) Act, 1972.<br />

Note: the scientific names are given as spelled in the Act – a number <strong>of</strong> spelling errors are present in the listing.<br />

1. Schedule 1. Part IV (Collection and trade, including gift, prohibited) Liphyra brassolis<br />

Allotinus drumila<br />

A. fabious penormis<br />

Amblopala avidiena<br />

Amblypodia ace arala<br />

A. alea constanceae<br />

A. ammonariel<br />

A. arvina ardea<br />

A. asopia<br />

A. comica<br />

A. opalina<br />

A. zeta<br />

Biduanda melisa cyana<br />

Callophyrs leechii<br />

Castalius rosimon alarbus<br />

Charana cepheis<br />

Chlioria othona<br />

Deudoryx epijarbas amatius<br />

Everes moorei<br />

Gerydus biggsii<br />

G. symethus diopeithes<br />

Heliophorus hybrida<br />

Horaga albimacula<br />

Jamides ferrari<br />

Darkie, crenulate/great<br />

Angled darkie<br />

Hairstreak, Chinese<br />

Leaf blue<br />

Rosy oakblue<br />

Malayan bush blue<br />

Purple brown tailless oakblue<br />

Plain tailless oakblue<br />

Comic oakblue<br />

Opal oakblue<br />

Andaman tailless oakblue<br />

Blue posy<br />

Hairstreak, ferruginous<br />

Pierrot, common<br />

Mandarin blue, Cachar<br />

Tit, orchid<br />

Cornelian, scarce<br />

Cupid, Moore's<br />

Bigg's brownie<br />

Great Brownie<br />

Sapphires<br />

Onyxes<br />

Caeruleans<br />

11<br />

Listeria dudgenni<br />

Logania watsoniana subfasciata<br />

Lycaenopsis binghami<br />

L. haraldus ananga<br />

L. puspa prominent<br />

L. quadriplaga dohertyi<br />

Nacaduba noreia hampsoni<br />

Polymatus orbitulus leela<br />

Pratapa icetas mishmia<br />

Simiskina phalena harterti<br />

Sinthusa virgo<br />

Spindasis elwesi<br />

S. rukmini<br />

Strymon mackwoodi<br />

Tajuria ister<br />

T. luculentus nela<br />

T. yajna yajna<br />

Thecla ataxua zulla<br />

T. bieti mendera<br />

T. letha<br />

T. paona<br />

T. pavo<br />

Virachala smilis<br />

Date <strong>of</strong> legislation<br />

(1980)<br />

(1979)<br />

(1986)<br />

(1985)<br />

(1986)<br />

(1986)<br />

(1986)<br />

(1982)<br />

(1982)<br />

(1980)<br />

(1986)<br />

(1980)<br />

(1986)<br />

(1986)<br />

(1986)<br />

(1986)<br />

(1980)<br />

(1986)<br />

(1980)<br />

Butterfly, moth<br />

Lister's hairstreak<br />

Mottle, Watson's<br />

Hedge blue<br />

Hedge blue, Felder's<br />

Common hedge blue<br />

Naga hedge blue<br />

Limeblue, white-tipped<br />

Greenish mountain blue<br />

Royal, dark blue<br />

Brilliant, broadbanded<br />

Spark, pale<br />

Silverline, Elwe's<br />

Silverline, khaki<br />

Hairstreak, Mackwood's<br />

Royal, uncertain<br />

Royal, Chinese<br />

Royal, chestnut and black<br />

Wonderful hairstreak<br />

Indian purple hairstreak<br />

Watson's hairstreak<br />

Paona hairstreak<br />

Peacock hairstreak<br />

Guava blues<br />

Continued...


Table 4 (cont.). <strong>Lycaenidae</strong> listed in The Indian Wildlife (Protection) Act, 1972.<br />

2. Schedule II. Part II. ('Special game':<br />

Allotinus subviolaceous manychus<br />

Amblypodia abetrans<br />

A. aenea<br />

A. agaba aurelia<br />

A. agrata<br />

A. alesia<br />

A. apidanus ahamus<br />

A. areste areste<br />

A. bazaloides<br />

A. camdeo<br />

A. ellisi<br />

A. fulla ignara<br />

A. ganesa watsoni<br />

A. paragenesa zephpreeta<br />

A. paralea<br />

A. silhetensis<br />

A. suffusa stiffusu<br />

A. yendava<br />

Apharitis lilacinus<br />

Araotes lapithis<br />

Artipe eryx<br />

Bindahara phocides<br />

Bolhrinia chennellia<br />

Castraleus roxus manluena<br />

Catapoecilma delicalum<br />

C. elegans myoslina<br />

Charana jalindra<br />

Cheritrella truncipennis<br />

Chliaria kina<br />

Deudoryx hypargyria gaetulia<br />

Enchrysops enejus<br />

Everes kala<br />

Helipphorus androcles moorei<br />

Horaga onyx<br />

H. viola<br />

Hypolycaena nilgirica<br />

H. ihecloides nicobarica<br />

lraota rochana boswelliana<br />

Jamides alectokandulana<br />

J. exleodus para<br />

J. coeruleus<br />

J. kankena<br />

Lampides boeticus<br />

Lilacea albocaerulea<br />

L. atroguttata<br />

L. lilacea<br />

L. melaena<br />

L. minima<br />

Logania massalia<br />

Lycaenesthes lycaenina<br />

Mahathala ameria<br />

M. atkinsoni<br />

Magisba malaya presbyter<br />

Nacaduba aluta coelestis<br />

N. ancyra aberrans<br />

N. dubiosa fulva<br />

N. helicon<br />

N. herus major<br />

N. pactolus<br />

Neucheritra febronia<br />

licence needed to collect or trade.) Niphanda cymbia<br />

Orthomiella ponlis<br />

Pithecops fulgens<br />

Polymmatus devanica devanica<br />

P. metallica metallica<br />

P. orbitulus jaloka<br />

P. younghusbandi<br />

Poritia erycinoides elisei<br />

P. hewitsoni<br />

P. plusrata geta<br />

Pralapa bhetes<br />

P. blanka<br />

P. deva<br />

P. icetas<br />

Rapala buxaria<br />

R. chandrana chandrana<br />

R. nasaka<br />

R. refulgens<br />

R. rubida<br />

R. scintilla<br />

R. sphinx sphinx<br />

R. varuna<br />

Spindasis elima elima<br />

S. lohita<br />

S. nipalicus<br />

Suasa lisidus<br />

Surendra todara<br />

Tajuria albiplaga<br />

T. cippus cippus<br />

T. culta<br />

T. diaeus<br />

T. illurgioides<br />

T. illurgis<br />

T. jangala andamanica<br />

T. melastigma<br />

T. sebonga<br />

T. thyia<br />

T. yajna istroides<br />

T. callinara<br />

Tarucus dharta<br />

Thaduka multicaudata kanara<br />

Thecla ataxus ataxus<br />

T. bitel<br />

T. icana<br />

T. jakamensis<br />

T. kabreea<br />

T. khasia<br />

T. kirbariensis<br />

T. suroia<br />

T. syla assamica<br />

T. vittata<br />

T. ziha<br />

T. zoa<br />

Una usta<br />

Yasoda tripunctata<br />

12<br />

3. Schedule IV. ('Small game':<br />

collect.)<br />

Tarucus ananda<br />

small game hunting licence needed to


Table 5. <strong>Lycaenidae</strong> included on the 1990 <strong>IUCN</strong> Red List <strong>of</strong> Threatened Animals.<br />

Taxon<br />

Alaena margaritacea<br />

Aloeides caledoni<br />

A. dentatis<br />

A. egerides<br />

A. lutescens<br />

Argyrocupha malagrida malagrida<br />

A. m. paarlensis<br />

Callophrys mossii bayensis<br />

Capys penningtoni<br />

Chrysoritis cotrelli<br />

C. oreas<br />

C. zeuxo<br />

Cyclyrius mandersi<br />

Deloneura immaculata<br />

D. millari millari<br />

Deudorix penningtoni<br />

D. vansoni<br />

Durbania limbata<br />

Erikssonia acraeina<br />

Eumaeus atala florida<br />

Everes comyntax texanus<br />

Glaucopsyche lygdamus palosverdesensis<br />

G. xerces<br />

Hemiargus thomasi bethune-bakeri<br />

Icaricia icarioides missionensis<br />

I. i. moroensis<br />

I. i. pheres<br />

Lepidochrysops ariadne<br />

L. bacchus<br />

L. hypolia<br />

L. loewensteini<br />

L. lotana<br />

L. methymna dicksoni<br />

L. titei<br />

Lycaeides argyrognomon lotis<br />

L. melissa samuelis<br />

Lycaena dispar<br />

L. dorcas claytoni<br />

L. hermes<br />

Maculinea alcon<br />

M. arion<br />

M. arionides<br />

Status<br />

V<br />

V<br />

R<br />

V<br />

v<br />

V<br />

V<br />

E<br />

E<br />

I<br />

I<br />

I<br />

I<br />

Ex?<br />

V<br />

V<br />

V<br />

V<br />

R<br />

V<br />

Ex<br />

Ex?<br />

Ex<br />

I<br />

E<br />

I<br />

I<br />

E<br />

V<br />

Ex?<br />

V<br />

E<br />

E<br />

V<br />

E<br />

I<br />

E<br />

I<br />

I<br />

V<br />

V<br />

V<br />

13<br />

Country<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

U.S.A.<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

Mauritius<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

U.S.A.<br />

(U.S.A.)<br />

U.S.A.<br />

(U.S.A.)<br />

U.S.A.<br />

U.S.A.<br />

U.S.A.<br />

U.S.A.<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

U.S.A.<br />

U.S.A.<br />

northern Europe<br />

U.S.A.<br />

Mexico, U.S.A.<br />

Europe, USSR<br />

Europe, USSR<br />

China, Japan, USSR<br />

Continued...


Table 5. (cont.) <strong>Lycaenidae</strong> included on the 1990 <strong>IUCN</strong> Red List <strong>of</strong> Threatened Animals.<br />

Taxon<br />

M. nausithous<br />

M. rebeli<br />

M. teleius<br />

M. t. burdigalensis<br />

Notarthrinus binghami<br />

Oreolyce dohertyi<br />

Ornipholidotos peucetia penningtoni<br />

Oxychaeta dicksoni<br />

Panchala ganesa loomisi<br />

Philotiella speciosa bohartorum<br />

Plebejus emigdionis<br />

P. icarioides missionensis<br />

Plebicula golgus<br />

Poecilmitis adonis<br />

P. aureus<br />

P. endymion<br />

P. lyncurium<br />

P. nigricans<br />

P. rileyi<br />

Pseudalmenus chlorinda chlorinda<br />

P. c. conara<br />

Pseudiolaus lulua<br />

Shijimiaeoides battoides allyni<br />

S. b. comstocki<br />

S. enoptes smithi<br />

S. lanstoni<br />

S. rita mattoni<br />

Spindasis collinsi<br />

Strymon acis bartrami<br />

S. avalona<br />

Thestor dicksoni dicksoni<br />

T. kaplani<br />

T. tempe<br />

Trimenia wallegrenii<br />

Uranothauma usambarae<br />

(Categories defined in Appendix 1).<br />

Status<br />

E<br />

V<br />

E<br />

E<br />

R<br />

R<br />

I<br />

E<br />

E<br />

I<br />

I<br />

E<br />

E<br />

V<br />

I<br />

V<br />

V<br />

V<br />

V<br />

I<br />

I<br />

V<br />

E<br />

I<br />

E<br />

I<br />

I<br />

V<br />

I<br />

K<br />

V<br />

V<br />

V<br />

V<br />

V<br />

14<br />

Country<br />

Europe, USSR<br />

south & central Europe<br />

Europe, northern Asia<br />

France<br />

India<br />

India<br />

Mozambique, S. Africa<br />

South Africa<br />

Japan<br />

U.S.A.<br />

U.S.A.<br />

U.S.A.<br />

Spain<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

Australia: Tasmania<br />

Australia: Tasmania<br />

South Africa<br />

U.S.A.<br />

U.S.A.<br />

U.S.A.<br />

U.S.A.<br />

U.S.A.<br />

Tanzania<br />

U.S.A.<br />

U.S.A.<br />

South Africa<br />

South Africa<br />

South Africa<br />

South Africa<br />

Tanzania


Increasing human recreational activities constitute another<br />

serious threat to many habitats all over the world and the<br />

following list, while far from exhaustive, gives some idea <strong>of</strong> the<br />

range <strong>of</strong> habitats involved:<br />

• coastal sand dunes in California are threatened by <strong>of</strong>f-road<br />

vehicles and trampling;<br />

• alpine heathlands and meadows in Europe and southeastern<br />

Australia are threatened by the construction <strong>of</strong> ski-lifts,<br />

runs, access roads, car parks and resort accommodation and<br />

facilities;<br />

• Pacific islands habitats are threatened by the proliferation <strong>of</strong><br />

golf courses and by the exotic vegetation <strong>of</strong>ten introduced;<br />

• mangrove swamps in eastern Australia are threatened by<br />

coastal resort development.<br />

In many cases these recreational activities involve the<br />

degradation <strong>of</strong> particularly sensitive habitats which tend to<br />

support isolated, relict and <strong>of</strong>ten taxonomically discrete<br />

populations <strong>of</strong> lycaenids and other insects.<br />

Examples could be multiplied several-fold, but the principle<br />

is well established, and the vital importance <strong>of</strong> suitable habitat<br />

and resources for conserving small animal and plant populations<br />

should not need further emphasis.<br />

Pollution<br />

The effects <strong>of</strong> chemical pollution on lycaenids are difficult to<br />

assess, but a number <strong>of</strong> declines <strong>of</strong> particular species in Europe<br />

have been attributed in part to atmospheric pollution, including<br />

acid rain. Such pollution is likely to affect the well being <strong>of</strong><br />

sensitive foodplants and a wide spectrum <strong>of</strong> invertebrates<br />

associated with vegetation. Likewise, pesticide drift may cause<br />

occasional hazard, both in agricultural and forest environments.<br />

Exotic introductions<br />

No significant information is available on the deleterious effects<br />

<strong>of</strong> exotic taxa on native <strong>Lycaenidae</strong> in many parts <strong>of</strong> the world.<br />

The introduction <strong>of</strong> Dutch Elm Disease into Britain, with<br />

consequent large scale demise <strong>of</strong> Ulmus trees, led to a reduction<br />

in the numbers <strong>of</strong> the Whiteletter Hairstreak, Strymonidia walbum<br />

(Knoch). Similarly, the introduction <strong>of</strong> myxomatosis to<br />

Britain in the 1950s resulted in a drastic reduction <strong>of</strong> the<br />

intensity <strong>of</strong> rabbit grazing on chalk grasslands, with a resultant<br />

reduction in the numbers <strong>of</strong> several butterfly species, such as<br />

Lysandra bellargus (Rottemburg) and Maculinea arion. In<br />

New Zealand, Gibbs (1980) noted the trend towards decline <strong>of</strong><br />

Zizina oxleyi (Felder & Felder) in parts <strong>of</strong> the South Island<br />

through hybridisation with the invasive Australian Z. labradus<br />

(both are sometimes treated as subspecies <strong>of</strong> Z. otis (F.) (Figure<br />

3)). Claims that hybridisation with the introduced Strymon<br />

melinus Hübner could threaten the Avalon Hairstreak, S. avalona<br />

(Wright), on Santa Catalina Island, California, need further<br />

investigation (Wells et al. 1983).<br />

In a rather different interaction between exotic and native<br />

species, Brown (1990) reported that the widespread Leptotes<br />

marina (Reakirt) had adapted well to urbanisation in North<br />

15<br />

America with its range expansion largely due to a 'switch' to a<br />

South African larval foodplant (Plumbago auriculata), which<br />

is used widely in freeway landscaping and as an ornamental.<br />

Furthermore, larvae associate closely with the introduced<br />

Argentine ant, and the prime nectar source for adults is a<br />

Brazilian tree. Leptotes thus has benefited from the presence <strong>of</strong><br />

several different exotic species.<br />

<strong>Conservation</strong> <strong>of</strong> <strong>Lycaenidae</strong><br />

The major sequence <strong>of</strong> needs in order to effect conservation<br />

programmes for any form <strong>of</strong> terrestrial wildlife is as follows:<br />

i) Documentation and education, to increase awareness at all<br />

levels and as a mode <strong>of</strong> communication between informed<br />

scientists and those who make practical decisions over<br />

priorities for land use;<br />

ii) Detection <strong>of</strong> habitats supporting either critical faunas or<br />

single notable or vulnerable species which merit protection<br />

and the promotion <strong>of</strong> their continued protection in existing<br />

National Parks and other reserves;<br />

iii) Investigation <strong>of</strong> the limits and/or wisdom <strong>of</strong> legislative<br />

protection for particular taxa or habitats, as an interim<br />

measure whilst additional documentation is obtained, and<br />

iv) Autecological studies <strong>of</strong> selected taxa as a basis for<br />

formulating sound management plans, a step which can<br />

come only from a basis <strong>of</strong> substantial research rather than<br />

haphazard extrapolation and which is, therefore, costly.<br />

v) Investigation <strong>of</strong> techniques for captive rearing, in case <strong>of</strong><br />

need for ex situ conservation, or translocation. This should<br />

not be seen as a replacement option for in situ conservation.<br />

The information contained in this volume has hitherto been<br />

scattered through a wide range <strong>of</strong> reports (<strong>of</strong> varying degrees <strong>of</strong><br />

formality and distribution) and scientific papers. In dealing<br />

with such a diverse group <strong>of</strong> insects, this book cannot be as<br />

definitive as 'Threatened Swallowtail <strong>Butterflies</strong>' (Collins and<br />

Morris 1985), but the examples given reflect a growing number<br />

<strong>of</strong> detailed studies on lycaenids and concern over their<br />

conservation. It represents a useful starting point for the<br />

development <strong>of</strong> conservation programmes for the <strong>Lycaenidae</strong><br />

and should help to focus attention on groups or geographical<br />

regions in need <strong>of</strong> urgent attention.<br />

Public awareness<br />

Most conservation projects to date have been species orientated<br />

and many <strong>of</strong> these have done much to improve public awareness<br />

<strong>of</strong> the threats to lycaenid species. The potential for<br />

reintroductions is demonstrated well by the recent project<br />

involving the successful liberation <strong>of</strong> the Large Blue, Maculinea<br />

arion L. in Britain. This case is discussed in detail by Elmes and<br />

Thomas (1992) and by New ('Large Blues', this volume).<br />

Attempts to introduce the continental subspecies <strong>of</strong> the Large<br />

Copper, Lycaena dispar, into Britain, are <strong>of</strong> long standing<br />

(Duffey 1977 and this volume). Both projects highlight the


Figure 3. Distribution <strong>of</strong> Zizina oxyleyi (native) and Z. labradus (exotic) in<br />

New Zealand, and their hybrid zone (after Gibbs 1980).<br />

need for substantial amounts <strong>of</strong> ecological information for such<br />

management initiatives and they have done much to improve<br />

public awareness <strong>of</strong> butterfly conservation issues.<br />

There is no doubt that particular 'charismatic' butterflies<br />

amongst the <strong>Lycaenidae</strong> can do much to increase public<br />

awareness <strong>of</strong> conservation at a level which is otherwise difficult<br />

to realise, as witness the Large Blue campaign in Britain, the<br />

Mission Blue (Plebejus icarioides missionensis Hovanitz) in<br />

California and the more recent Eltham Copper (Paralucia<br />

pyrodiscus lucida Crosby) campaign in Australia.<br />

However, such species-orientated conservation is essentially<br />

confined to well-developed countries. It could not be achieved<br />

practically in much <strong>of</strong> the rest <strong>of</strong> the world, where conservation<br />

<strong>of</strong> any sort is considered a luxury and where there are few<br />

skilled entomological or conservationist practitioners and little<br />

if any local funding or sympathy for insect well-being.<br />

Any opportunity for promotion <strong>of</strong> significant species as<br />

' local emblems' is worth pursuing. For example the appearance<br />

<strong>of</strong> butterflies on stamps might help to gain public sympathy:<br />

lycaenids have been depicted on postage stamps <strong>of</strong> nearly 70<br />

countries (Coles et al. 1991 and Table 6) and these have<br />

included both less developed and developed countries.<br />

16<br />

Identification <strong>of</strong> threatened species and critical<br />

faunas<br />

The identification <strong>of</strong> these is clearly an immediate need in<br />

lycaenid conservation. The information contained in this volume<br />

indicates that while the former is the main focus in the developed<br />

world it is the latter which may be the most relevant path to<br />

follow elsewhere. The largest group <strong>of</strong> oriental Polyommatini<br />

(the 'Lycaenopsis group') was appraised by Eliot and Kawazoe<br />

(1983), and additional comprehensive studies <strong>of</strong> this sort would<br />

be <strong>of</strong> considerable value in setting conservation priorities on a<br />

faunistic basis. That study indicated a number <strong>of</strong> regions with<br />

high levels <strong>of</strong> diversity and/or endemism which could thus be<br />

considered as critical faunas (see Collins and Morris 1985 on<br />

Papilionidae, Ackery and Vane-Wright 1984 on Nymphalidae:<br />

Danainae).<br />

Other examples <strong>of</strong> critical faunas can be cited: the Philippines<br />

support more species <strong>of</strong> the Polyommatini than any other area<br />

<strong>of</strong> comparable size and some are endemic; Sulawesi is less rich<br />

than the Philippines but about half the species there are relict<br />

endemics or have only a narrow range elsewhere; most taxa in<br />

the Papuan subregion are endemic; centres <strong>of</strong> diversity (?refugia)<br />

for Riodininae occur in the neotropics (see taxa accounts for<br />

South America, this volume) and <strong>of</strong> other taxa in various parts<br />

<strong>of</strong> the world. Many <strong>of</strong> these areas are undergoing substantial,<br />

rapid and irreversible changes at present, and the prime need is<br />

to increase reservation <strong>of</strong> ecologically representative areas as a<br />

prelude to more detailed studies <strong>of</strong> their needs. Unless additonal<br />

reserves are forthcoming the rich diversity <strong>of</strong> Riodininae in the<br />

neotropics (see essays by Brown and Callaghan, this volume)<br />

and the substantial numbers <strong>of</strong> unusual Polyommatinae in<br />

southeast Asia will be seriously reduced.<br />

The emphasis on species-orientated lycaenid conservation<br />

is likely to occupy entomologists in temperate regions for some<br />

time to come. This has two important results. Firstly, detailed<br />

studies <strong>of</strong> particular taxa, which would be impossible in a<br />

broader context, may give a good basis for their proper<br />

management. Various projects on the reintroduction <strong>of</strong> species<br />

and successful management <strong>of</strong> threatened species, which are<br />

documented in the taxa accounts, attest to the depth <strong>of</strong> ecological<br />

knowledge necessary for the successful management <strong>of</strong> habitat<br />

for a particular species. In fact, for most <strong>Lycaenidae</strong>, preservation<br />

<strong>of</strong> suitable habitat is generally not in itself sufficient to ensure<br />

their long-term well-being: intricate management plans to<br />

conserve complex tripartite associations involving butterfly,<br />

ant and plant and to control seral succession by ensuring that<br />

some rapidly developing early stages may be continually<br />

available, are integral facets <strong>of</strong> successful conservation.<br />

Secondly, particular critical habitats may be reserved for the<br />

species and give benefit also to less obvious rare biota. However,<br />

few such studies result in reservation <strong>of</strong> large or well-buffered<br />

habitats. Systems akin to the British SSSI (Sites <strong>of</strong> Special<br />

Scientific Interest, history documented by Moore 1987) are not<br />

yet widespread elsewhere but merit serious attention in other<br />

parts <strong>of</strong> the world to ensure the adequate reservation <strong>of</strong> wellbuffered<br />

habitats.


Table 6. List <strong>of</strong> <strong>Lycaenidae</strong> which have been depicted on postage stamps up to October 1991. (Data from Coles et al. 1991; identifications not checked<br />

further, authors' names omitted).<br />

Species<br />

Aethiopana honorius<br />

Citrinophila erastus<br />

Hewitsonia boisduvali<br />

Mimacraea marshalli<br />

Pentila abraxas<br />

Telipna acraea<br />

Ogyris amaryllis<br />

Aphnaeus questiauxi<br />

Aphniolaus pallene<br />

Atlides polybe<br />

Axiocerses amanga<br />

A. harpax<br />

A. styx<br />

Bindahara phocides<br />

Callophrys crethona<br />

Chrysozephyrus mushaellus<br />

Deudorix epijarbas<br />

Epamera handmani<br />

E. sidus<br />

Eumaeus alula<br />

Evenus coronata<br />

E. dindymus<br />

E. regalis<br />

Hypokopelates otraeda<br />

Hypolycaena antifaunus<br />

Japonica lutea<br />

Lipaphnaeus leonina<br />

Loxura atymnus<br />

Myrina silenus<br />

Narathura centaurus<br />

Panthiades bathildis<br />

Pratapa ctesia<br />

Pseudalmenus chlorinda<br />

Pseudolycaena marsyas<br />

Rapala arata<br />

Ritra aurea<br />

Scoptes alphaeus<br />

Spindasis modesla<br />

S. natalensis<br />

S. nyassae<br />

Strymon maesites<br />

S. martialis<br />

S. melinus<br />

S. ruf<strong>of</strong>usca<br />

S. simaethis<br />

Stugeta marmorea<br />

Tanuetheira timon<br />

Thecla betulae<br />

Agrodiaetus amanda<br />

A. damon<br />

Country, year <strong>of</strong> issue<br />

Ghana 1990<br />

Ghana 1990<br />

Uganda 1989<br />

Uganda 1989<br />

Ghana 1990<br />

Ghana 1990<br />

Australia 1981<br />

Zambia 1980<br />

Uganda 1989<br />

Grenada 1975<br />

Uganda 1989<br />

Mozambique 1953, Togo 1990<br />

Tanzania 1973<br />

Palau 1990<br />

Jamaica 1978<br />

China 1963<br />

Samoa 1986<br />

Malawi 1966<br />

Kenya 1988<br />

Cuba 1974<br />

Ecuador 1970, Nicaragua 1986<br />

Grenadines <strong>of</strong> St. Vincent 1975<br />

Belize 1990, Brazil 1979, Nicaragua<br />

1967<br />

Mali 1964<br />

Togo 1990<br />

Laos 1986<br />

Mali 1964<br />

China 1963<br />

Angola 1982, Mauritania 1966, Togo<br />

1990. Sierra Leone 1979, Tanzania 1988<br />

Malaysia 1970<br />

Belize 1974<br />

Laos 1986<br />

Australia 1981<br />

Grenada 1990, Grenadines <strong>of</strong> Grenada<br />

1985, St. Vincent 1978<br />

Korea (North) 1977<br />

Nicaragua 1986<br />

South Africa 1977<br />

Zambia 1980<br />

Swaziland 1987<br />

Malawi 1984<br />

Dominica 1988, Turks & Caicos Islands<br />

1982, Grenadines <strong>of</strong> St. Vincent 1989<br />

Cayman Islands 1988<br />

Honduras 1991<br />

Grenadines <strong>of</strong> Grenada 1985<br />

Grenada 1989, Grenadines <strong>of</strong> Grenada<br />

1985<br />

Sierra Leone 1987<br />

Congo 1971, Sierra Leone 1987<br />

Bulgaria 1990<br />

Finland 1990<br />

Mongolia 1963<br />

17<br />

Species<br />

Aricia agestis<br />

Brephidium exilis<br />

Catochrysops taitensis<br />

Celastrina argiolus<br />

Cupidopsis jobates<br />

Danis danis<br />

Glaucopsyche melanops<br />

Hemiargus ammon<br />

H. hanno<br />

H. thomasi<br />

Jamides bochus<br />

J. cephion<br />

Lampides boelicus<br />

Leptotes cassius<br />

Luthrodes cleotas<br />

Lycaeides idas<br />

Lysandra albicans<br />

L. beltargus<br />

L. coridon<br />

Maculinea arion<br />

Meleageria daphnis<br />

Polyommatus icarus<br />

Tarucus balkanicus<br />

Uranothauma crawshayi<br />

Heliphorus epicles<br />

Heodes solskyi<br />

H. virgaureae<br />

Lycaena dispar<br />

L. helle<br />

L. salustius<br />

Palaeochrysophanus hippothoe<br />

Thersamonia phoebus<br />

T. thersamon<br />

Abisara talantus<br />

A ncyluris formosissima<br />

A. jurgenseni<br />

Chorinea faunus<br />

Dodona adonira<br />

Helicopis cupido<br />

Melanis pixe<br />

Nymphidium mantus<br />

Nymula orestes<br />

Rhetus thia<br />

R. sp.<br />

Stalachtis calliope<br />

S. phlegia<br />

Country, year <strong>of</strong> issue<br />

Cyprus 1983<br />

Cayman Islands 1990, Turks & Caicos<br />

Islands 1990<br />

Samoa 1986<br />

Libya 1981<br />

Togo 1990<br />

Netherlands New Guinea 1960<br />

Cyprus 1983, Libya 1981<br />

Cayman Islands 1988, Cuba 1991<br />

Barbados 1983, British Virgin Islands<br />

1978<br />

Turks & Caicos Islands 1990<br />

Tonga 1989<br />

Solomon Islands 1980<br />

Ascension 1987, Fiji 1985, Vanuatu<br />

1991<br />

St. Vincent 1989<br />

Vanuatu 1983, Wallis & Futuna Islands<br />

1987<br />

Canada 1988<br />

Libya 1981<br />

Hungary 1959<br />

Switzerland 1952<br />

Gibraltar 1977, Great Britain 1981,<br />

Hungary 1969, Poland 1967<br />

Bulgaria 1962, Hungary 1966, Rumania<br />

1969<br />

Guernsey 1981, Ireland 1 985, Malta<br />

1986<br />

Turkey 1958<br />

Malawi 1973<br />

Hong Kong 1979, Macau 1985<br />

China 1963<br />

Finland 1990, Hungary 1959, Mongolia<br />

1977, Rumania 1960<br />

Germany (F.R.) 1991<br />

Germany (F.R.) 1991<br />

New Zealand 1970<br />

Hungary 1974, Nicaragua 1986<br />

Ifni 1963<br />

Hungary 1969<br />

Sierra Leone 1987<br />

Guinea 1973, Hungary 1984<br />

Nicaragua 1967<br />

Guyana 1989<br />

China 1963<br />

Guyana 1983, Surinam 1971<br />

Nicaragua 1967<br />

Guyana 1983<br />

Grenada 1975<br />

Nicaragua 1967<br />

Honduras 1991<br />

Guyana 1989, Surinam 1971<br />

Surinam 1971


Managing lycaenid populations<br />

Various management plans for particular <strong>Lycaenidae</strong> have<br />

been produced, and these have several elements in common:<br />

habitat security; the need for management based on sound<br />

biological and ecological understanding <strong>of</strong> the target species;<br />

and the need for an understanding <strong>of</strong> the endangering processes.<br />

Increasing public awareness <strong>of</strong> the particular lycaenid is also<br />

commonly seen as important. The processes needed are well<br />

exemplified in a 'flow-chart' produced by Arnold (1983a) for<br />

Smith's Blue (Euphilotesenoptessmithi(Mattoni)) in California<br />

and augmented in his Recovery Plans for the Lotis Blue<br />

(Lycaeides idas lotis (Lintner) and other species (USFWS<br />

1984, 1985), and this has been used as a basis for the scheme<br />

outlined in Table 7. It is clear from this that the basic information<br />

necessary to formulate sound management for any ecologically<br />

sensitive species cannot be gathered instantly. All too <strong>of</strong>ten the<br />

time available is insufficient once a decision has been taken to<br />

develop a habitat containing a threatened species or population.<br />

The approach <strong>of</strong> 'Population Viability Analysis' (sometimes,<br />

'Population Vulnerability Analysis'), PVA, is receiving<br />

substantial attention in assessing conservation needs <strong>of</strong> vertebrate<br />

animals. In general, the detailed demographic and reproductive<br />

data needed for such predictive modelling are not available for<br />

extending this approach to invertebrates. However, a North<br />

American lycaenid, the Karner Blue (Lycaeidesmelissa samuelis<br />

Nabokov) has recently (April 1992) been the focus <strong>of</strong> the first<br />

specialist workshop held to appraise an insect in this context.<br />

This species lives in fire-successional vegetation, and the<br />

metapopulations are associated with local extinctions and<br />

recolonisation or habitat shifts linked with climate and fire<br />

regimes (Cushman and Murphy, this volume). The outcome<br />

from this workshop may mark a substantial augmentation to<br />

current methods <strong>of</strong> formulating management programmes for<br />

insects <strong>of</strong> conservation concern.<br />

Ex situ conservation<br />

Table 7. A pro-forma scheme for species-orientated conservation <strong>of</strong> <strong>Lycaenidae</strong> (after Arnold 1983a).<br />

1. Preserve, protect and manage known existing habitat to provide conditions needed by the species.<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

Preserve: prevent further degradation, development or environmental modification.<br />

Steps (some or all):<br />

(i) Cooperative agreements with landowners/managers.<br />

(ii) Memoranda or undertakings.<br />

(iii) <strong>Conservation</strong> easements.<br />

(iv) Site acquisition (purchase/donation): private land.<br />

(v) Site reservation: public land.<br />

Maintain larval and adult resources at known habitat(s).<br />

Steps (some or all):<br />

(i) Minimise use <strong>of</strong> toxic substances: herbicides, pesticides.<br />

(ii) Minimise uncontrolled intrusion by humans: trampling, <strong>of</strong>f-road vehicles, etc.<br />

(iii) Minimise intrusion by domestic stock: cattle, horses, sheep, etc.<br />

(iv) Minimise exotic vegetation planting.<br />

(v) Minimise removal <strong>of</strong> native vegetation, unless controlled.<br />

The controlled harvesting and ranching <strong>of</strong> Papilionidae to<br />

reduce pressure on natural populations and supply collector<br />

needs has proved to be an important conservation avenue for<br />

this group. It probably will never be achievable (or, indeed,<br />

needed) for lycaenids on any large scale, despite their uniqueness<br />

and vulnerability and their geographical overlap with areas<br />

where Papilionidae are also vulnerable. However, rearing <strong>of</strong><br />

particular rare species in captivity does have some potential for<br />

augmenting field populations and for effecting translocations.<br />

The Atala Hairstreak, Eumaeus atala, has proved to be relatively<br />

easy to rear in captivity, and this practice is pursued by several<br />

butterfly houses in Britain using stock from Florida (Collins<br />

1987a). Some success has been obtained with attempts to start<br />

new colonies <strong>of</strong> E. atala in the wild, and although the species<br />

continues to be regarded as vulnerable to habitat loss it has<br />

become more widespread in recent years (Lenczewski 1980).<br />

Some Japanese Theclini may be reared relatively easily from<br />

eggs collected in the field, and eggs <strong>of</strong> species which are<br />

otherwise very hard to collect have been found (Kuzuya 1959).<br />

The larval stage lasts for 3–4 weeks, as does the pupal period,<br />

and the only major problem in rearing was a tendency for older<br />

larvae to be cannibalistic.<br />

Unlike most other butterflies, captive rearing <strong>of</strong> lycaenids<br />

may need to incorporate provision for ants, <strong>of</strong>ten <strong>of</strong> particular<br />

species for any given lycaenid, as well as foodplants.<br />

Propose critical habitat (U.S.).<br />

If necessary, clarify taxonomic status <strong>of</strong> target lycaenid in habitat and, where known, outlier or other populations.<br />

18<br />

Continued...


Table 7 (cont.). A pro-forma scheme for species-orientated conservation <strong>of</strong> <strong>Lycaenidae</strong> (after Arnold 1983a).<br />

2. Manage and enhance lycaenid population(s) by habitat maintenance and quality improvement, and reducing effects <strong>of</strong> limiting factors.<br />

(a) Investigate and initiate habitat improvement methods as appropriate.<br />

Examples:<br />

(i) Remove or control exotic weedy or noxious plants.<br />

(ii) Promote natural establishment <strong>of</strong> foodplants and other natural vegetation – if necessary, propagate and transplant,<br />

(iii) Promote particular grazing (grassland) or coppicing (woodland) regimes.<br />

(b) Determine physical and climatic regimes/factors needed by species and relate to local habitat enhancement in overall site.<br />

(c) Investigate ecology <strong>of</strong> the lycaenid species<br />

References<br />

(i) Life history and phenology; dependence on particular plant species/stages/organs.<br />

(ii) Dependence on ants or other animals, and their role in protection from predators and parasites.<br />

(iii) Population status; size, movement, degree <strong>of</strong> isolation, sex ratio, etc.<br />

(iv) Adult behaviour: mating, oviposition cues and sites, activity rhythms,<br />

(v) Determine predators, parasitoids and other factors which cause mortality or limit population growth.<br />

(vi) Investigate possibility <strong>of</strong> captive rearing from local population enhancement or range extension.<br />

(d) Investigate ecology <strong>of</strong> tending ant species and (if needed) homoptcrous prey.<br />

(e) Investigate ecology <strong>of</strong> food plant species<br />

(i) Life history and recruitment processes.<br />

(ii) Mortality and debilitatory factors, including other consumer species.<br />

(iii) Limiting factors – edaphic conditions, slope, exposure, etc.<br />

(iv) If needed, horticultural studies to determine propagation techniques for transplantation/augmentation <strong>of</strong> food plant stocks.<br />

3. Evaluate above, and incorporate into development <strong>of</strong> long-term management plan. Computer modelling may assist in making management<br />

decisions.<br />

4. Monitor lycaenid populations to determine their status and to evaluate success <strong>of</strong> management.<br />

(a) Determine sites to be surveyed, if choice available and/or logistics limited.<br />

(b) Develop methodology to estimate population numbers, distribution and trends in abundance.<br />

Examples:<br />

(i) Counts <strong>of</strong> larvae when feeding at night.<br />

(ii) Counts <strong>of</strong> adults by transect patrols, mark-recapture, etc.<br />

5. Throughout all above, increasing public awareness <strong>of</strong> the species by education/information programmes.<br />

Examples:<br />

(i) Information signs at key sites (unless risk <strong>of</strong> inducing unwanted intrusion, collecting, etc.)<br />

(ii) Interpretive tours, if secure sites permit without causing additional problems.<br />

(iii) Audio and visual programmes, publications.<br />

– TV/radio interviews and information on the species and its management.<br />

– <strong>Conservation</strong> education programmes for schools and community groups.<br />

– 'Popular style' newspaper articles.<br />

– Continuing liaison with all interested parties.<br />

6. Enforce available regulations and laws to protect species. Determine whether additional legal steps needed, and promote these if necessary.<br />

ACKERY, P.R. 1984. Systematic and faunistic studies on butterflies. In:<br />

Vane-Wright, R.I. and Ackery, P.R. (Eds). The <strong>Biology</strong> <strong>of</strong> <strong>Butterflies</strong>.<br />

Academic Press, London, pp. 9–21.<br />

ACKERY, P.R. and VANE-WRIGHT, R.I. 1984. Milkweed <strong>Butterflies</strong>. British<br />

Museum (Natural History, London and Cornell University Press, New<br />

York.<br />

ARNOLD, R.A. 1983a. <strong>Conservation</strong> and management <strong>of</strong> the endangered<br />

Smith's blue butterfly, Euphilotes enoptes smithi (Lepidoptera:<br />

<strong>Lycaenidae</strong>).J. Res. tepid. 22: 135–153.<br />

19<br />

ARNOLD, R.A. 1983b. Ecological studies <strong>of</strong> six endangered butterflies:<br />

island biogeography, patch dynamics and the design <strong>of</strong> nature reserves.<br />

Univ. Calif. Publns Ent. 99: 1–161.<br />

AUSTIN, G.T. 1972. A possible case <strong>of</strong> mimicry between lycaenid butterflies<br />

(<strong>Lycaenidae</strong>). J. tepid. Soc. 26: 63–64.<br />

BAYLIS, M. and PIERCE, N.E. 1991. The effect <strong>of</strong> host-plant quality on the<br />

survival <strong>of</strong> larvae and oviposition by adults <strong>of</strong> an ant-attended lycaenid<br />

butterfly, Jalmenus evagoras. Ecol. Entomol. 16: 1–9.<br />

BOWERS, M.D. and LARIN, Z. 1989. Acquired chemical defence in the<br />

lycaenid butterfly, Eumaeus atala. J. Chem. Ecol. 15: 1133–1146.<br />

BREEDLOVE, D.E. and EHRLICH, P.R. 1968. Plant-herbivores coevolution:<br />

lupines and lycaenids. Science 162: 672–673.


BREEDLOVE, D.E. and EHRLICH, P.R. 1972. Coevolution: patterns <strong>of</strong><br />

legume predation by a lycaenid butterfly. Oecologia 10: 99–104.<br />

BRIDGES, C.A. 1988. Catalogue <strong>of</strong> <strong>Lycaenidae</strong> and Riodinidae(Lepidoptera:<br />

Rhopalocera). Urbana, Illinois.<br />

BROWN, J.W. 1990. Urban biology <strong>of</strong> Leptotes marina (Reakirt)(<strong>Lycaenidae</strong>).<br />

J. tepid. Soc. 44: 200–201.<br />

BROWN, R.M. and OPLER, P.A. 1967. Biological observations <strong>of</strong> Callophrys<br />

viridis (<strong>Lycaenidae</strong>). J. Lepid. Soc. 21: 113–114.<br />

CALLAGHAN, C.J. 1979. A new genus and a new subspecies <strong>of</strong> Riodinidae<br />

from southern Brazil. Bull. Allyn Mus. 53: 1–7.<br />

CALLAGHAN, C.J. 1983. A study <strong>of</strong> isolating mechanisms among neotropical<br />

butterflies <strong>of</strong> the subfamily Riodininae. J. Res. Lepid. 21: 159–176.<br />

CALLAGHAN, C.J. 1986. Notes on the biology <strong>of</strong> Stalachtis susanna<br />

(<strong>Lycaenidae</strong>: Riodininae) with a discussion <strong>of</strong> riodinine larval strategies.<br />

J. Res. Lepid. 24: 258–263.<br />

CLENCH, H.K. 1955. Revised classification <strong>of</strong> the butterfly family <strong>Lycaenidae</strong><br />

and its allies. Ann. Carneg. Mus. 33: 261–274.<br />

CLENCH, H.K. 1966. Behavioral thermoregulation in butterflies. Ecology<br />

47: 1021–1034.<br />

CLENCH, H.K. and MILLER, L.D. 1980. Papilio ladon Cramer vs. Argus<br />

pseudargiolus Boisduval & Le Conte (<strong>Lycaenidae</strong>): a nomenclatorial<br />

nightmare. J. Lepid. Soc. 34: 103–118.<br />

COLES, A., PHIPPS, T. and other Members <strong>of</strong> the Butterfly and Moth Stamp<br />

Society. 1991. Collect <strong>Butterflies</strong> and other Insects on Stamps. Stanley<br />

Gibbons, London and Ringwood.<br />

COLLINS, N.M. 1987a. Butterfly houses in Britain, the conservation<br />

implications. <strong>IUCN</strong>, Cambridge.<br />

COLLINS, N.M. 1987b. Legislation to conserve insects in Europe. Pamphlet<br />

13. Amateur Entomologists' Society, London.<br />

COLLINS, N.M. and MORRIS, M.G. 1985. Threatened Swallowtail <strong>Butterflies</strong><br />

<strong>of</strong> the World. <strong>IUCN</strong>, Gland and Cambridge.<br />

COTTRELL, C.B. 1984. Aphytophagy in butterflies: its relationship to<br />

myrmecophily. Zool. J. Linn. Soc. 79: 1–57.<br />

CROAT, T.B. 1978. Flora <strong>of</strong> Barro Colorado Island. Stanford University<br />

Press, Stanford, California.<br />

DE VRIES, P.J. 1988. The use <strong>of</strong> epiphylls as larval hostplants by the<br />

neotropical riodinid butterfly, Sarota gyas. J. nat. Hist. 22: 1447–1450.<br />

DE VRIES, P.J. 1990. Enhancement <strong>of</strong> symbiosis between butterfly caterpillars<br />

and ants by vibrational communication. Science 248: 1104–1106.<br />

DE VRIES, P.J. 1991 a. Evolutionary and ecological patterns in myrmecophilous<br />

riodinid butterflies. In: Huxley, C.R. and Cutler, D.F. (Eds) Ant-Plant<br />

Interactions. Oxford University Press, Oxford, pp.143–156.<br />

DE VRIES, P.J. 1991b. Mutualism between Thisbe irenea butterflies and ants,<br />

and the role <strong>of</strong> ant ecology in the evolution <strong>of</strong> larval-ant associations. Biol.<br />

J. Linn. Soc. 43: 179–195.<br />

DOWNEY, J.C. 1962. Host-plant relations as data for butterfly classification.<br />

Syst. Zool. 11: 150–159.<br />

DOWNEY, J.C. 1965. Mimicry and distribution <strong>of</strong> Caenurgina caerulea Grt.<br />

(Noctuidae). J. Lepid. Soc. 19: 165–170.<br />

DOWNEY, J.C. 1966. Sound production in pupae <strong>of</strong> <strong>Lycaenidae</strong>. J. Lepid.<br />

Soc. 20: 129–155.<br />

DOWNEY, J.C. and FULLER, W.C. 1961. Variation in Plebejus icarioides.<br />

I. Foodplant specificity. J. Lepid. Soc. 15: 34–42.<br />

DUFFEY, E. 1977. The re-establishment <strong>of</strong> the large copper butterfly Lycaena<br />

dispar batavus on Woodwalton Fen NNR, Cambridgeshire, England<br />

1969–1973. Biol. Conserv. 12: 143–158.<br />

EDWARDS, W.H. 1886. On the history and preparatory stages <strong>of</strong> Feniseca<br />

tarquinis Fabr. Can. Ent. 18: 141–153.<br />

EHRLICH, P.R. 1958. The comparative morphology, phylogeny and higher<br />

classification <strong>of</strong> the butterflies. Univ. Kans. Sci. Bull. 39: 305–370.<br />

EHRLICH, P.R. and EHRLICH, A. 1972. Wing-shape and adult resources in<br />

lycaenids. J. Lepid. Soc. 26: 196–197.<br />

ELIOT, J.N. 1973. The higher classification <strong>of</strong> the <strong>Lycaenidae</strong>: a tentative<br />

arrangement. Bull. Brit. Mus. Nat. Hist. (Ent.) 28: 373–506.<br />

ELIOT, J.N. and KAWAZOE, A. 1983. Blue butterflies <strong>of</strong> the Lycaenopsis<br />

group. British Museum (Natural History), London.<br />

ELMES, G.W. and THOMAS, J.A. 1992. Complexity <strong>of</strong> species conservation<br />

20<br />

in managed habitats: interaction between Maculinea butterflies and their<br />

ant hosts. Biodiv. and Conserv. 1: 155–169.<br />

FIEDLER, K. 1991a. European and North West African <strong>Lycaenidae</strong><br />

(Lepidoptera) and their associations with ants. J. Res. Lepid. 28: 239–257<br />

(1989).<br />

FIEDLER, K. 1991b. Systematic, evolutionary and ecological implications <strong>of</strong><br />

myrmecophily within the <strong>Lycaenidae</strong>. Bonn, Zool. Monograph 21.<br />

FIEDLER, K. and MASCHWITZ, U. 1988. Functional analysis <strong>of</strong> the<br />

myrmecophilous relationships between ants (Hymenoptera: Formicidae)<br />

and lycaenids (Lepidoptera: <strong>Lycaenidae</strong>). II. Lycaenid larvae as trophobiotic<br />

partners <strong>of</strong> ants – a quantitative approach. Oecologia 75: 204–206.<br />

FIEDLER, K. and MASCHWITZ, U. 1989. The symbiosis between the<br />

weaver ant, Oecophylla smaragdina, and Anthene emolus, an obligate<br />

myrmecophilous lycaenid butterfly. J. nat. Hist. 23: 833–846.<br />

FUJII, H. 1989. Repeated copulation in an orange hairstreak, Shirozua janasi:<br />

a case <strong>of</strong> mate guarding? J. Lepid. Soc. 43: 68–71.<br />

GIBBS, G.W. 1980. New Zealand <strong>Butterflies</strong>. Collins, Auckland.<br />

GOODPASTURE, C. 1974. Foodplant specificity in the Plebejus (Icaricia)<br />

acmon group <strong>Lycaenidae</strong>). J. Lepid. Soc. 28: 53–63.<br />

HARVEY, D.J. 1987. The higher classification <strong>of</strong> the Riodinidae (Lepidoptera).<br />

Ph.D. Thesis, University <strong>of</strong> Texas, Austin. (not seen: referred to by De<br />

Vries, 1991a).<br />

HEATH, J. 1981. Threatened Rhopalocera in Europe. Council <strong>of</strong> Europe,<br />

Strasbourg.<br />

HENNING, S.F. 1983. Biological groups within the <strong>Lycaenidae</strong> (Lepidoptera).<br />

J. ent. Soc. sth. Afr. 46: 65–85.<br />

HINTON, H.E. 1951. Myrmecophilous <strong>Lycaenidae</strong> and other Lepidoptera –<br />

a summary. Proc. Trans. S. Lond. ent. nat. Hist. Soc. 1949–50: 111–175.<br />

HINTON, H.E. 1974. Lycaenid pupae that mimic anthropoid heads. J. Entomol.<br />

(A) 49: 65–69.<br />

<strong>IUCN</strong>, 1990. The <strong>IUCN</strong> Red List <strong>of</strong> Threatened Animals. Gland and Cambridge.<br />

IWASE, T. 1953. Aberrant feeders among Japanese lycaenid larvae. Lepid.<br />

News 7: 45–46.<br />

JOHNSON, K. 1984. An attempted interfamilial mating (<strong>Lycaenidae</strong>-<br />

Nymphalidae).J. Res. Lepid. 28: 291–292.<br />

JOHNSON, K. and BORGO, P.M. 1976. Patterned perching behaviour in two<br />

Callophrys (Mitoura) (<strong>Lycaenidae</strong>). J. Lepid. Soc. 30: 169–183.<br />

KITCHING, R.L. 1981. Egg clustering and the southern hemisphere lycaenids:<br />

comments on a paper by N.E. Stamp. Amer. Nat. 118: 423–425.<br />

KITCHING, R.L. 1987. Aspects <strong>of</strong> the natural history <strong>of</strong> the lycaenid butterfly<br />

Allotinus major in Sulawesi. J. nat. Hist. 21: 535–544.<br />

KUZUYA, T. 1959. The breeding <strong>of</strong> the Theclini and collecting their eggs in<br />

winter.J. Lepid. Soc. 13: 175–181.<br />

LENCZEWSKI, B. 1980. <strong>Butterflies</strong> <strong>of</strong> Everglades National Park. Report<br />

T–588. National Parks Service, Florida.<br />

LUNDGREN, L. 1977. The role <strong>of</strong> intra- and interspecific male: male<br />

interactions in Polyommatus icarus Rott. and some other species <strong>of</strong> blues<br />

(<strong>Lycaenidae</strong>). J. Res. Lepid. 16: 249–264.<br />

MACNEILL, CD. 1967. A unidirectional mass movement by Satyrium<br />

saepium (<strong>Lycaenidae</strong>). J. Lepid. Soc. 21: 204.<br />

MALICKY, H. 1969. Versuch einer Analyse der ökologischen Beziehungen<br />

zwischen Lycaeniden und Formiciden. Tidschr. Ent. 112: 213–298.<br />

MOORE, N.W. 1987. The Bird <strong>of</strong> Time. Cambridge University Press,<br />

Cambridge.<br />

MURPHY, D.D. 1989. Are we studying our endangered butterflies to death?<br />

7. Res. Lepid. 26: 236–239.<br />

NAKAMURA, I. 1976. Female anal hair tuft in Nordmannia myratle: eggcamouflaging<br />

function and taxonomic significance. J. Lepid. Soc. 30:<br />

305–309.<br />

NEW, T.R., BUSH, M.B., THORNTON, I.W.B. and SUDARMAN, H.K.<br />

1988. The butterfly fauna <strong>of</strong> the Krakatau islands after a century <strong>of</strong><br />

colonisation. Phil. Trans. R. Soc. Lond. B. 322: 445–457.<br />

OWEN, D.F. 1991. Pseudaletis leonis: a rare mimetic butterfly in a West<br />

African rainforest (Lepidoptera: <strong>Lycaenidae</strong>). Tropical Lepidoptera 2:<br />

111–113.<br />

PIERCE, N.E. 1983. The ecology and evolution <strong>of</strong> symbioses between lycaenid<br />

butterflies and ants. Ph.D. Thesis, Harvard University.


PIERCE, N.E. 1984. Amplified species diversity: case study <strong>of</strong> an Australian<br />

lycaenid butterfly and its attendant ants. In: Vane-Wright, R.I. and Ackery,<br />

P.R. (Eds) The <strong>Biology</strong> <strong>of</strong> <strong>Butterflies</strong>. Academic Press, London, pp.<br />

197–200.<br />

PIERCE, N.E. 1985. Lycaenid butterflies and ants: selection for nitrogenfixing<br />

and other protein-rich foodplants. Amer. Nat. 125: 888–895.<br />

PIERCE, N.E. 1987. The evolution and biogeography <strong>of</strong> associations between<br />

lycaenid butterflies and ants. Oxford Surv. evol. Biol. 4: 89–116.<br />

PIERCE, N.E. and EASTEAL, S.I. 1986. The selective advantage <strong>of</strong> attendant<br />

ants for the larvae <strong>of</strong> a lycaenid butterfly Glaucopsyche lygdamus. J. anim.<br />

Ecol. 55: 451–462.<br />

PIERCE, N.E. and MEAD, P.S. 1981. Parasitoids as selective agents in the<br />

symbiosis between butterfly larvae and ants. Science 211: 1185–1187.<br />

PRATT, G.F. and BALLMER, G.R. 1991. Acceptance <strong>of</strong> Lotus scoparius<br />

(Fabaceae) by larvae <strong>of</strong> <strong>Lycaenidae</strong>. J. Lepid. Soc. 45: 188–196.<br />

RIOTTE, J.C. and UCHIDA, G. 1979. <strong>Butterflies</strong> <strong>of</strong> the Hawaiian islands<br />

according to the stand <strong>of</strong> late 1976. J. Res. Lepid. 17: 33–39.<br />

ROBBINS, R.K. 1980. The lycaenid 'false head' hypothesis: historical review<br />

and quantitative analysis.J. Lepid. Soc. 34: 194–208.<br />

ROBBINS, R.K. 1981. The lycaenid 'false head' hypothesis: predation and<br />

wing pattern variation <strong>of</strong> lycaenid butterflies. Amer. Nat. 118: 770–775.<br />

ROBBINS, R.K. 1982. How many butterfly species? News Lepid. Soc. 1982:<br />

40–41.<br />

ROBBINS, R.K. and AIELLO, A. 1982. Foodplant and oviposition records<br />

for Panamanian <strong>Lycaenidae</strong> and Riodinidae. J. Lepid. Soc. 36: 65–75.<br />

ROBBINS, R.K. and SMALL, G.B. 1981. Wind dispersal <strong>of</strong> Panamanian<br />

hairstreak butterflies and its evolutionary significance. Biotropica 13:<br />

308–315.<br />

ROBINSON, R. 1971. Lepidoptera genetics. Pergamon, Oxford.<br />

SANDS, D.P.A. 1986. A revision <strong>of</strong> the genus Hypochrysops C. & R. Felder<br />

(Lepidoptera: <strong>Lycaenidae</strong>). Entomonograph No. 7. Brill, Leiden.<br />

SCOTT, J.A. 1968. Hilltopping as a mating mechanism to aid the survival <strong>of</strong><br />

low density species. J. Res. Lepid. 7: 191–204.<br />

SCOTT, J.A. 1973. Down-valley flights <strong>of</strong> adult Theclini in search <strong>of</strong><br />

21<br />

nourishment. J. tepid. Soc. 27: 283–287.<br />

SHAPIRO, A.M. 1982. An interfamilial courtship (<strong>Lycaenidae</strong>, Pieridae). J.<br />

Res. tepid. 20 (1981): 54.<br />

SHAPIRO, A.M. 1985. An intersubfamilial courtship (<strong>Lycaenidae</strong>). J. Res.<br />

tepid. 24: 195.<br />

SHIELDS, O. 1967. Hilltopping. J. Res. tepid. 6: 69–178.<br />

SHIELDS, O.1989. World numbers <strong>of</strong> butterflies.J, tepid. Soc. 43:178–183.<br />

SHIROZU, T. and YAMAMOTO, H. 1957. Systematic position <strong>of</strong> the genus<br />

Curetis (Lepidoptera: Rhopalocera). Sieboldia 2: 43–51.<br />

SIBATANI, A. 1984. A remarkable polymorphism <strong>of</strong> mature larvae <strong>of</strong> Zizina<br />

labradus (Godart), common grass blue butterfly (Lepidoptera: <strong>Lycaenidae</strong>)<br />

from the Sydney area. Aust. ent. Mag. 11: 21–26.<br />

SIBATANI, A. 1992. Observations on the period <strong>of</strong> active flight in males <strong>of</strong><br />

Favonius (<strong>Lycaenidae</strong>) in southern Primor'e, the Russian Federation. Tyô<br />

to Ga 43: 23–34.<br />

STEMPFFER, H. 1957. Les lepidoptères de l'Afrique noire francaise (3):<br />

Lycaenides. Init. afr. 14: 1–228.<br />

STEMPFFER, H. 1967. The genera <strong>of</strong> the African <strong>Lycaenidae</strong> (Lepidoptera:<br />

Rhopalocera). Bull. Br. Mus. nat. Hist. (Ent.). Supplement 10, 322 pp.<br />

THOMAS, J.A., MUNGUIRA, M.L., MARTIN, J. and ELLIS, G.W. 1991.<br />

Basal hatching by Maculinea butterfly eggs: a consequence <strong>of</strong> advanced<br />

myrmecophily? Biol. J. Linn. Soc. 44: 175–184.<br />

USFWS (U.S. Fish and Wildlife Service) 1984. Recovery plan for the San<br />

Bruno Elfin and Mission Blue butterflies. U.S. Fish and Wildlife Service,<br />

Portland, Oregon.<br />

USFWS (U.S. Fish and Wildlife Service) 1985. Recovery plan for the Lotis<br />

Blue butterfly. U.S. Fish and Wildlife Service, Portland, Oregon.<br />

VALENTINE, P.S. and JOHNSON, S.J. 1989. Polyphagy in larvae <strong>of</strong><br />

Hypochrysops miskini (Waterhouse) (Lepidoptera: <strong>Lycaenidae</strong>). Aust.<br />

ent. Mag. 16: 1–3.<br />

WELLS, S.M., PYLE, R.M. and COLLINS, N.M. 1983. The <strong>IUCN</strong> Invertebrate<br />

Red Data Book. <strong>IUCN</strong>, Gland.<br />

ZIEGLER, J.B. and ESCALANTE, T. 1964. Observations on the life history<br />

<strong>of</strong> Callophrys xami (<strong>Lycaenidae</strong>). J. Lepid. Soc. 18: 85–89.


PART 2. REGIONAL ASSESSMENTS<br />

This section is important both for what it contains and for what<br />

it does not contain. It demonstrates that most <strong>of</strong> our knowledge<br />

<strong>of</strong> the status <strong>of</strong> particular <strong>Lycaenidae</strong> is derived from temperate<br />

regions, particularly in the northern hemisphere, and that the<br />

fauna <strong>of</strong> much <strong>of</strong> the Old World tropics (in particular) cannot<br />

yet be evaluated in this way. Much <strong>of</strong> the tropics supports a high<br />

diversity <strong>of</strong> butterflies, but few resident lepidopterists with the<br />

facilities for undertaking biological studies: much <strong>of</strong> what can<br />

be inferred is based less on ecological surveys and a detailed<br />

knowledge <strong>of</strong> biology and more on old collections and synoptic<br />

works. For example, books such as those by Corbet and<br />

Pendlebury (1978, Malaysia) and Seki et al. (1991, Borneo)<br />

give invaluable appraisals <strong>of</strong> those magnificent local faunas.<br />

The introduction to the latter includes passionate comment on<br />

conservation <strong>of</strong> butterflies in Borneo, and the text, photographs<br />

and keys facilitate identification <strong>of</strong> 379 species and subspecies<br />

– but the status <strong>of</strong> many <strong>of</strong> these is largely unknown, and the<br />

knowledge <strong>of</strong> their distribution <strong>of</strong>ten fragmentary. Many<br />

lycaenids there, as elsewhere, are indeed both rare and restricted<br />

– perhaps threatened by habitat alteration or destruction — but<br />

Introductory comment<br />

22<br />

it is not possible to give a sound biological overview <strong>of</strong> these<br />

unique faunas, except in such very general terms.<br />

There is abundant need to document lycaenid diversity in<br />

many <strong>of</strong> the nominal 'protected areas' in the tropics, and to<br />

ensure that the widest possible range <strong>of</strong> habitats is conserved,<br />

especially for the many specialist forest-frequenting taxa.<br />

Though few specific details are available, there seems little<br />

doubt that continuing forest destruction throughout the African<br />

and Asian tropics is adversely affecting these insects.<br />

References<br />

CORBET, A.S. and PENDLEBURY, H.M. 1978. The <strong>Butterflies</strong> <strong>of</strong> the Malay<br />

Peninsula. 3rd ed., revised by Eliot, J.N., Malayan Nature Society, Kuala<br />

Lumpur.<br />

SEKI, Y., TAKANAMI, Y. and OTSUKA, K. 1991. <strong>Butterflies</strong> <strong>of</strong> Borneo<br />

2(1). <strong>Lycaenidae</strong>. Tobishima Corporation, Tokyo.


<strong>Conservation</strong> biology <strong>of</strong> <strong>Lycaenidae</strong>: A European overview<br />

Miguel L. MUNGUIRA 1 , José MARTIN 1 and Emilio BALLETTO 2<br />

1 Departamento de Biología, Universidad Autónoma de Madrid, Cantoblanco, 28049 Madrid, Spain<br />

2 Dipartimento di Biologia Animate, Università di Torino, Via Accademia Albertina 17, 10123 Torino, Italy<br />

<strong>Biology</strong> <strong>of</strong> European <strong>Lycaenidae</strong><br />

The biology <strong>of</strong> European lycaenids is generally well known as<br />

a result <strong>of</strong> both general and specific studies. A review <strong>of</strong> central<br />

European species was made by Malicky (1969a) who dealt with<br />

larval foodplants, phenology and overwintering stages.<br />

Relationships with ants were also reviewed by Malicky (1969b)<br />

and Fiedler (1990a, 1991). Books on regional faunas including<br />

detailed ecological information on a species-by-species basis<br />

are available for Italy (Verity 1943), the Netherlands (Tax<br />

1989), British Isles (Ford 1970; Emmet and Heath 1989) and<br />

Switzerland (SBN 1987). Specific autecological studies cover<br />

a wide range <strong>of</strong> species from almost every major group within<br />

the European fauna. Some <strong>of</strong> these studies are listed in Table 1,<br />

together with the geographic areas in which they were<br />

undertaken. Several studies in press or short notices have not<br />

been listed and a wider range <strong>of</strong> species is covered by these.<br />

Thus the information available covers roughly 50% <strong>of</strong> the<br />

European lycaenids. The Table clearly shows wider coverage<br />

<strong>of</strong> northern species and <strong>of</strong> the countries where the present<br />

authors are based (Italy and Spain), but nevertheless it gives a<br />

general idea <strong>of</strong> our knowledge <strong>of</strong> European lycaenids.<br />

While most <strong>of</strong> the northern species are well known as far as<br />

their biology is concerned, information on most Mediterranean<br />

species is not available, even on such basic topics as foodplants<br />

or habitat preferences. Greece, with its 12 endemic species, is<br />

the area where basic studies are most urgently needed. High<br />

altitude species are also under-represented in ecological studies<br />

due to the difficulties <strong>of</strong> reaching their habitats for long-term<br />

studies, but their conservation seems to pose fewer problems<br />

than that <strong>of</strong> lowland species.<br />

Synecological studies are generally less abundant and only<br />

refer to some geographic areas such as southern France (Cléu<br />

1950; Bigot 1952, 1956; Dufay 1961, 1965–66), Hungary<br />

(Uherkovic 1972,1975,1976), Czechoslovakia (Kralichek and<br />

Povolny 1978), Italy (Balletto et al. 1977, 1982a–e, 1985,<br />

1988, 1989), Switzerland (Erhardt 1985) and Germany<br />

(Kratochwil 1989a, 1989b, 1989c), or to some particular<br />

ecosystems (Fouassin 1961; Janmoulle 1965; Cléu 1957;<br />

Mikkola and Spitzer 1983; see also Ehrlich 1984).<br />

23<br />

Larval foodplants<br />

Most <strong>of</strong> the European lycaenids are polyommatines (Kudrna<br />

1986). The data on lycaenid larval foodplants in Table 2 clearly<br />

illustrate the fact that polyommatines have radiated in our area<br />

more than the other groups, using wider niches and evolving to<br />

produce a wider range <strong>of</strong> endemic species and peculiar ecotypes.<br />

Life cycles<br />

The most common strategy adopted by the European lycaenids<br />

to endure adverse seasons is larval quiescence. Most<br />

polyommatines hide away and spend the winter season (in<br />

northern Europe) or the summer and winter (in the Mediterranean<br />

areas) as second or third instar larvae. When the egg or pupa is<br />

the overwintering stage, a diapause normally takes place (e.g.<br />

Iolana iolas (Ochsenheimer), Tomares ballus (F): Munguira<br />

1989, Jordano et al. 1990). Theclines typically overwinter as<br />

eggs, lycaenines and polyommatines as larvae. In the last case,<br />

however, there is considerable variation, again supporting the<br />

idea <strong>of</strong> a wider radiation <strong>of</strong> the group in our area. Table 3<br />

summarises the data on overwintering stages for Spanish<br />

lycaenids (a sample covering 70% <strong>of</strong> European species).<br />

There is one generation per year for most species. This<br />

seems to be a clear adaptation to temperate climates which have<br />

a favourable summer and a cold winter in which insect life<br />

comes through a dormant period. All the theclines follow this<br />

pattern, but again the polyommatines show some variation,<br />

having species with two generations (e.g. Polyommatus [or<br />

Lysandra] bellargus (Rottemburg), Lycaeides idas (L.)) or as<br />

many generations as weather allows (e.g. Tarucus theophrastus<br />

(F.), Aricia cramera Eschscholtz, Polyommatus icarus<br />

(Rottemburg)). It is possible that species with two generations<br />

are not 'fixed' to this condition, but constrained by food<br />

availability or unfavourable weather conditions. Thus, many<br />

species with a low number <strong>of</strong> generations may increase the<br />

number when resource limits change. An example <strong>of</strong> this is<br />

Polyommatus icarus with one generation in northern Britain,<br />

two in southern Britain (Emmet and Heath 1989) and four to<br />

five in Spain (Martin 1982).


Table 1. European Lycaenid species which have been the subject <strong>of</strong> thorough autecological studies, and the areas where these studies were made.<br />

Nomenclature here, and in the text follows Kudrna (1986).<br />

Species<br />

Thecla betulae<br />

Satyrium pruni<br />

S. ilicis<br />

Quercusia quercus<br />

Callophrys rubi<br />

Tomares ballus<br />

Lycaena virgaureae<br />

L. phlaeas<br />

L. dispar<br />

Lampides boeticus<br />

Leptotes pirithous<br />

Glaucopsyche alexis<br />

G. melanops<br />

Maculinea rebeli<br />

M. alcon<br />

M. arion<br />

M. nausithous<br />

M. teleius<br />

Cupido minimus<br />

C. lorquinii<br />

Iolana iolas<br />

Cyaniris semiargus<br />

Plebejus argus<br />

P. pylaon<br />

Lycaeides idas<br />

Aricia morronensis<br />

A. agestis<br />

A. artaxerxes<br />

A. cramera<br />

A. nicias<br />

A. eumedon<br />

Polyommatus golgus<br />

P. dorylas<br />

P. nivescens<br />

P. thersites<br />

P. bellargus<br />

P. albicans<br />

P. hispana<br />

P. humedasae<br />

P. galloi<br />

P. icarus<br />

Agriades glandon<br />

A. zullichi<br />

Area<br />

U.K.<br />

U.K.<br />

Italy<br />

U.K.<br />

Germany<br />

Spain<br />

Sweden<br />

U.K.<br />

U.K.<br />

Spain<br />

Spain<br />

Spain<br />

Spain<br />

France<br />

Spain<br />

France<br />

Spain<br />

U.K.<br />

France<br />

France<br />

U.K.<br />

Spain<br />

Spain<br />

Spain<br />

U.K.<br />

Germany<br />

Spain<br />

Hungary<br />

Austria<br />

Sweden<br />

Belgium<br />

Spain<br />

U.K.<br />

U.K.<br />

Denmark<br />

Italy<br />

Sweden<br />

Spain<br />

Spain<br />

Spain<br />

Spain<br />

France<br />

U.K.<br />

France<br />

Italy<br />

Italy<br />

Spain<br />

U.K.<br />

Spain<br />

Spain<br />

24<br />

Reference<br />

Thomas 1974<br />

Thomas 1974<br />

Fiori 1957<br />

Thomas 1975<br />

Fiedler 1990b<br />

Jordano et al. 1990<br />

Douwes 1975<br />

Dempster 1971; Emmet and Heath 1989<br />

Duffey 1968<br />

Martin 1984<br />

Martin 1984<br />

Martin 1981<br />

Martin 1981<br />

Thomas et al. 1989<br />

Munguira 1989<br />

Thomas et al. 1989<br />

Munguira 1989<br />

Thomas 1980<br />

Thomas 1984a<br />

Thomas 1984a<br />

Morton 1985<br />

Munguira 1989<br />

Munguira 1989<br />

Rodriguez 1991<br />

Thomas 1985; Ravenscr<strong>of</strong>t 1990<br />

Weidemann 1986<br />

Munguira 1989<br />

Bálint and Kertész 1990<br />

Malicky 1961<br />

Pellmyr 1983<br />

Leestmans 1984<br />

Munguira and Martin 1988<br />

Jarvis 1958, 1959; Hoegh-Guldberg and Jarvis 1970<br />

Hoegh-Guldberg and Jarvis 1970<br />

Balletto et al. 1981<br />

Wiklund 1977<br />

Munguira et al. 1988<br />

Munguira and Martin 1989<br />

Munguira and Martin 1989<br />

Munguira and Martin 1989<br />

Chapman 1914<br />

Davis et al. 1958; Thomas 1983<br />

Sourès 197;, Nel 1978<br />

Schurian 1980<br />

Manino et al. 1987<br />

Balletto and Toso 1979<br />

Martin 1984<br />

Dennis 1984<br />

Munguira 1989<br />

Munguira 1989


Table 2. Larval foodplants in the European <strong>Lycaenidae</strong>.<br />

Subfamily<br />

Polyommatinae<br />

Theclinae<br />

Lycaeninae<br />

Theclinae<br />

Lycaeninae<br />

Polyommatinae<br />

TOTAL<br />

Taxa<br />

Most feed mainly on legumes<br />

Some genera have shifted to other food plants, e.g.<br />

Tarucus Moore, Celastrina Tutt<br />

Some Maculinea Ecke; Pseudophilotes Beuret<br />

Maculinea arion (L.)<br />

Freyeria Courvoisier<br />

Scolitantides Hubner<br />

Plebejus argus (L.)<br />

Vacciniina Tutt<br />

Aricia Reichenbach<br />

Agriades Hubner<br />

Cyaniris Dalman<br />

Most species feed mainly on trees and shrubs<br />

Some are adapted to legumes, e.g.<br />

Tomares ballus (F.), Callophrys rubi (L.)<br />

All species are restricted to:<br />

Egg<br />

9 (75%)<br />

1 (17%)<br />

4 ( 8%)<br />

14 (20%)<br />

Food Plants<br />

Fabaceae<br />

Rhamnaceae, Aquifoliaceae, Cornaceae, Araliaceae<br />

Gentianaceae, Rosaceae, Lamiaceae<br />

Boraginaceae<br />

Crassulaceae<br />

Fabaceae, Ericaceae, Cistaceae<br />

Ericaceae<br />

Geraniaceae, Cistaceae,<br />

Primulaceae<br />

Fabaceae, Plumbaginaceae<br />

Fagaceae, Rhamnaceae, Oleaceae, Ulmaceae, Ericaceae<br />

Fabaceae<br />

Polygonaceae (Rumex and Polygonum spp.)<br />

Table 3. Frequencies and percentages <strong>of</strong> species overwintering at different stages in the three subfamilies <strong>of</strong> Spanish <strong>Lycaenidae</strong>. Data taken from Martin<br />

(1982).<br />

European habitats <strong>of</strong> importance for<br />

lycaenids<br />

Although conservation practice involves a battle for every<br />

piece <strong>of</strong> land containing wild animals or plants, some ecosystems<br />

in Europe are particularly important because they have become<br />

rare or support valuable species or communities (Blab and<br />

Kudrna 1982; Kudrna 1986; Balletto and Casale 1991). The<br />

following are some <strong>of</strong> the habitats <strong>of</strong> special importance for<br />

lycaenid conservation.<br />

25<br />

Larvae<br />

_<br />

5 (83%)<br />

40 (78%)<br />

45 (65%)<br />

Woodlands<br />

Pupae<br />

3 (25%)<br />

–<br />

7(14%)<br />

10(15%)<br />

Apart from regions <strong>of</strong> the extreme north <strong>of</strong> Europe, woodlands<br />

represent the only stable ecosystems from sea level up to the<br />

high altitude vegetational stages. Nevertheless they have been<br />

subject to the most severe human pressures, and very few<br />

remain in their natural condition.<br />

In the case <strong>of</strong> lycaenids, woodlands are generally used only<br />

by theclines which are mostly encountered on the borders <strong>of</strong><br />

natural or artificial clearings. Whether this is the natural<br />

condition, or if theclines (as inhabitants <strong>of</strong> the canopy) can<br />

rarely be observed in closed woods, remains for the moment a<br />

matter <strong>of</strong> speculation.


Woodland damage has been particularly severe in central<br />

Europe (e.g. Germany), but the wide distribution <strong>of</strong> most<br />

theclines makes threats to whole individual species rather<br />

unlikely. Only some species such as Satyrium pruni (L.), S.<br />

acaciae (F.), Thecla betulae (L.), or Callophrys avis Chapman<br />

are locally rare or have declined in the last few decades (Heath<br />

1981; Heath et al. 1984).<br />

Screes<br />

Rocky areas with sparse vegetation usually have a very interesting<br />

flora and fauna in Mediterranean countries. Limestone outcrops<br />

and schists are particularly interesting habitats for some rare<br />

endemics such as Aricia morronensis Ribbe. The most interesting<br />

areas are high altitude habitats where vegetation is sparse and the<br />

climate makes living conditions particularly severe for butterflies:<br />

here only a few well adapted species can survive. Some <strong>of</strong> the<br />

species living on these areas are among the rarest European<br />

endemics: Agriades zullichi Hemming and Polyommatus golgus<br />

(Hübner). Turanana panagaea (Herrich-Schaeffer), a rather<br />

widespread xerophilous species in Turkey, becomes extremely<br />

scarce and localised in Europe and although no appropriate<br />

habitat study has been made, it is possible that it may qualify as<br />

a member <strong>of</strong> this group on the western side <strong>of</strong> the Aegean.<br />

Grasslands<br />

Ecologically, grasslands can be subdivided into two major<br />

types on the basis <strong>of</strong> their being climacic or seral. The former<br />

occur typically in the far north <strong>of</strong> Europe and at high elevations<br />

<strong>of</strong> mountain ranges. The latter are found over a wider range <strong>of</strong><br />

altitudes and generally represent the direct or indirect result <strong>of</strong><br />

human activities. Within these grasslands some areas defined as<br />

'seminatural' have conserved a very rich soil and diverse<br />

vegetation and are by far the most important habitats for<br />

lycaenid conservation.<br />

Grassland management is changing dramatically throughout<br />

Europe, and the effects <strong>of</strong> this upon butterflies are just beginning<br />

to be understood. Erhardt and Thomas (1991) and Balletto et al.<br />

(1989) have dealt with the effect <strong>of</strong> grassland management on<br />

butterfly populations, stating that traditional land uses are the<br />

best way to preserve the existing butterfly diversity and are<br />

essential for the survival <strong>of</strong> several endangered species.<br />

As far as butterfly conservation is concerned several types<br />

<strong>of</strong> grasslands can be considered although these categories are<br />

very heterogeneous from the vegetational point <strong>of</strong> view.<br />

Dry grasslands and steppes are only found in Europe in some<br />

Mediterranean countries and in central east Europe. They <strong>of</strong>ten<br />

develop on stony, poor soils from the Mediterranean to the<br />

montane vegetation stage and are generally characterised by a<br />

comparatively low turf, or in some cases by the presence <strong>of</strong> one<br />

or more grass species <strong>of</strong> the genus Stipa from which some <strong>of</strong><br />

them derive their name.<br />

The xeric character <strong>of</strong> this type <strong>of</strong> vegetation is reflected in<br />

the structure <strong>of</strong> their butterfly communities which are dominated<br />

26<br />

by grass feeders (Satyrinae and Hesperiidae). Some more or<br />

less restricted endemics and other interesting lycaenids occur in<br />

these habitats, the most notable <strong>of</strong> these species being<br />

undoubtedly Plebejus pylaon Fischer von Waldheim. In<br />

mountain areas these grasslands host many Polyommatus species<br />

<strong>of</strong> the subgenus Agrodiaetus, some <strong>of</strong> which are extremely<br />

local.<br />

The conservation <strong>of</strong> these habitats is particularly important<br />

because traditionally they have been considered useless, and<br />

steppe landscapes are generally unattractive for the public and<br />

therefore given low priority for conservation. At low elevations,<br />

however, they have been used for sheep or goat grazing, mainly<br />

during the winter, or locally (e.g. Alps) as vineyards.<br />

Mesophilous grasslands represent the lusher, more humid<br />

counterpart <strong>of</strong> dry grasslands. They generally develop on deeper,<br />

richer soils and are very rich in lycaenid butterflies. The rare<br />

species present in such areas include Lycaena hippothoe (L.),<br />

Maculinea rebeli (Hirschke), M. arion, Aricia eumedon,<br />

Pseudophilotes bavius (Eversmann), Scolitantides orion (Pallas)<br />

and Lycaenides argyrognomon (Bergsträsser).<br />

Mountain meadows are maintained by extensive cattle or<br />

sheep grazing, or locally by seasonal mowing. As a consequence<br />

<strong>of</strong> the milk surplus in the EEC this management is changing to<br />

reafforestation for paper production, particularly in southern<br />

countries. On far too many occasions this management practice<br />

leads to plantations with alien tree species and to the total<br />

destruction <strong>of</strong> the original habitat.<br />

It is important to point out that the greatest diversity <strong>of</strong><br />

European lycaenids occurs in this kind <strong>of</strong> habitat, due to the<br />

ecotone nature <strong>of</strong> most meadows in which patches <strong>of</strong> trees,<br />

shrubs and grasslands occur in the same places. The beautiful,<br />

flowered meadows blooming with butterflies are disappearing<br />

very quickly in Europe and only some mountain areas such as<br />

the Alps or the Pyrenees act as refuge areas. The role <strong>of</strong><br />

extensive grazing is crucial to conserve these habitats, and<br />

subsidies will probably soon be needed to support some <strong>of</strong> the<br />

less productive farming practices. Other aggressive uses related<br />

to skiing and the expansion <strong>of</strong> mountain tourism are deleterious<br />

for the conservation <strong>of</strong> these habitats.<br />

High altitude grasslands are only present above the natural<br />

tree line, either at the highest elevations on the major mountains<br />

or in the extreme north <strong>of</strong> the continent. Their climacic character<br />

makes management less necessary and thus their conservation<br />

easier. An important number <strong>of</strong> the European rare lycaenids<br />

live in these habitats: all the Agriades species (including the<br />

extremely rare A. zullichi which is found in a habitat transitional<br />

with high altitude screes); Albulina orbitulus (Prunner), Cyaniris<br />

helena (Staudinger); and some subspecies <strong>of</strong> Aricia morronensis,<br />

Polyommatus eros (Ochsenheimer), P. eroides Firvaldszky,<br />

and P. golgus.<br />

Threats to this type <strong>of</strong> environment are comparatively<br />

limited and are mainly represented by the spreading <strong>of</strong> ski<br />

resorts and mountain tourism, or to extensive overgrazing in<br />

southwest Europe.


Wet meadows and grasslands are azonal, like screes, in the<br />

sense that they can develop almost at any altitude where<br />

appropriate soil conditions are met (in this case a high humidity).<br />

Particularly at low elevations they are among the most<br />

endangered habitats in Europe. Species living on these grasslands<br />

are threatened throughout their range and examples <strong>of</strong> these are<br />

some <strong>of</strong> the most endangered species <strong>of</strong> Maculinea: M. alcon<br />

(Denis and Schiffermüller); M. teleius (Bergsträsser); and M.<br />

nausithous (Bergsträsser). Threats to this type <strong>of</strong> habitat are<br />

<strong>of</strong>ten related to land drainage and include changes in land<br />

management as a consequence <strong>of</strong> the very productive nature <strong>of</strong><br />

these grasslands.<br />

Shrublands<br />

Shrublands normally grow as a consequence <strong>of</strong> the action <strong>of</strong><br />

disturbing ecological factors such as strong winds, short<br />

vegetation times, human management or recurrent fires. They<br />

are interesting because some rare species are almost restricted<br />

to these formations.<br />

True heathlands dominated by Erica species are the habitat <strong>of</strong><br />

Plebejus argus, a species that has declined rapidly in England<br />

but which does not seem to have similar conservation problems<br />

in other countries.<br />

Mediterranean shrublands (garigues or chaparral formations)<br />

also support interesting rare species such as Iolana iolas, a<br />

circum-Mediterranean species living on chaparral. Another<br />

peculiar type <strong>of</strong> Mediterranean shrubland is represented by<br />

heathlands (with Erica arborea and E. scoparia) found on the<br />

highest, windswept elevations <strong>of</strong> Corsica, Sardinia and Elba<br />

(France and Italy, Balletto et al. 1989). This is the exclusive<br />

habitat <strong>of</strong> the endemic Lycaeides Corsica (Tutt).<br />

Kretania psylorita (Freyer) is a rare endemic living in<br />

Cretan shrublands dominated by an Astragalus species (Leigheb<br />

et al. 1990). Similar to these are some species living on biotopes<br />

dominated by Thymus species (called locally in Spain<br />

'tomillares') where some endemics (Cupido lorquinii Herrich-<br />

Schaeffer) or rare species (Pseudophilotes bavius, P.<br />

abencerragus (Pierret) and Scolitantides orion (Pallas)) live.<br />

Subalpine-type shrublands are found on mountains above the<br />

tree line. They are characterised by the presence <strong>of</strong><br />

Rhododendron and Vaccinium, <strong>of</strong>ten in association with dwarf<br />

pine trees. Their butterflies are mainly boreal-alpine elements<br />

which include, in wet areas, Aricia nicias Meigen, and Vacciniina<br />

optilete (Knoch) among the lycaenids.<br />

Wetlands<br />

These are probably the most endangered among European<br />

habitats as a consequence <strong>of</strong> drainage either to control mosquitoes<br />

or to transform the habitat into agricultural land (particularly<br />

rice fields). Wetland drainage has affected huge areas, both in<br />

central Europe (see Kudrna 1986 for the case <strong>of</strong> Bohemia in the<br />

27<br />

16th century) and in the south <strong>of</strong> the continent (Italy in the 19th<br />

century). Typical lycaenids living on wetlands are Lycaena<br />

dispar (Haworth), L. helle (Denis and Schiffermüller) and at<br />

higher altitudes or latitudes, Vaciniina optilete. These three<br />

species are listed as threatened by Heath (1981) and L. dispar<br />

is listed in the Berne Convention as being one <strong>of</strong> the most<br />

endangered European lycaenids.<br />

Causes <strong>of</strong> decline and extinction <strong>of</strong><br />

European lycaenids<br />

This subject has been considered by several authors when<br />

dealing with the broader topics <strong>of</strong> European butterflies (Heath<br />

1981, Thomas 1984b, Kudrna 1986) and insects <strong>of</strong> the<br />

Mediterranean Basin (Balletto and Casale 1991). A considerable<br />

amount <strong>of</strong> information and a number <strong>of</strong> opinions are also<br />

available for individual countries (Blab and Kudrna 1982;<br />

Viedma 1984; Balletto and Kudrna 1985; SBN 1987; Gonseth<br />

1987; Swaay 1990; Bàlint 1991; Kulfan and Kulfan 1991).<br />

Although no single review has dealt with the subject solely with<br />

reference to lycaenids, many papers pointing out causes <strong>of</strong><br />

decline <strong>of</strong> individual species are now available (see Table 1 for<br />

references). An overview <strong>of</strong> the topic is provided in Table 4,<br />

where information on butterfly conservation at a European<br />

level is summarised.<br />

Habitat alteration or destruction<br />

All three authors referred to in Table 4, as well as various<br />

national reports, identify this factor as the most important one<br />

for butterfly decline throughout Europe. A change in habitat<br />

quality is the cause <strong>of</strong> all extinctions documented to have taken<br />

place. This applies also to the lycaenids Lycaena dispar and<br />

Maculinea arion in the United Kingdom (Duffey 1968; Thomas<br />

1980) and Lycaena hippothoe (L.), M. arion, M. nausithous and<br />

M. teleius in The Netherlands (Heath 1981).<br />

Some examples <strong>of</strong> documented extinctions in Europe are<br />

given in Table 5. Wetland and grassland destruction or alteration<br />

are the main causes <strong>of</strong> recent extinctions.<br />

Butterfly declines are reaching an alarming scale in most<br />

central and north European countries where a high proportion<br />

<strong>of</strong> the fauna is experiencing dramatic range reductions (Heath<br />

et al. 1984). The range reductions <strong>of</strong> the following lycaenid<br />

species have been caused by habitat changes: Plebejus argus<br />

(United Kingdom, Ravenscr<strong>of</strong>t 1990); Polyommatus bellargus<br />

(United Kingdom, Thomas 1983); Satyrium pruni (United<br />

Kingdom, Thomas 1974);LycaenadisparandL.helle(Germany,<br />

Kudrna 1986); and Polyommatus exuberans Verity and L.<br />

dispar (Italy, Balletto et al. 1982a–e, Balletto in press).<br />

Habitat alterations can be quite subtle: for example, a slight<br />

change <strong>of</strong> growth in grass height on British Maculinea arion<br />

sites is enough to make the habitat unsuitable for the butterfly<br />

host ant (Thomas 1989). This change produced by grazing<br />

relaxation was sufficient to cause the disappearance <strong>of</strong> the


Table 4. Causes <strong>of</strong> decline or extinction <strong>of</strong> European butterflies in three different reviews <strong>of</strong> the topic. Similar causes (abbreviated) are listed on the<br />

same line.<br />

Heath 1981<br />

• Habitat destruction<br />

wetland drainage, changes in<br />

grassland management,<br />

forestry practices<br />

• Air pollution<br />

• Pesticides<br />

• Climatic change<br />

• Urbanisation<br />

• Tourism<br />

• Collecting<br />

• Commerce<br />

Thomas 1984b<br />

• Habitat changes<br />

wetlands, woodlands,<br />

agricultural habitats<br />

• Forestry management<br />

• Air pollution<br />

• Insecticides<br />

• Weather & climate<br />

• Butterfly collectors<br />

• Isolation and area<br />

Kudrna 1986<br />

• Wetland drainage<br />

• Grassland management<br />

• Afforestation<br />

• Air pollution<br />

• Weed & pest control<br />

• Urbanisation<br />

• Tourism & transport<br />

• Overcollecting<br />

• Earthworks<br />

Table 5. Extinctions <strong>of</strong> European lycaenids documented to have taken place in individual countries with an indication <strong>of</strong> the total number <strong>of</strong> extinct<br />

butterfly species in each case.<br />

Country<br />

United Kingdom<br />

Netherlands<br />

Denmark<br />

Luxembourg<br />

Extinct<br />

lycaenids<br />

3<br />

5<br />

1<br />

3<br />

Species<br />

butterfly from formerly suitable habitats. Extensive grazing is<br />

declining rapidly over almost all <strong>of</strong> Europe due to intensification<br />

<strong>of</strong> agricultural practices and abandonment <strong>of</strong> former grazing<br />

areas. Neither grazing relaxation nor the decline in extensive<br />

grazing is good news for butterflies because these insects have<br />

evolved in Europe over thousands <strong>of</strong> years together with man,<br />

and have probably benefited from the patchy structure <strong>of</strong><br />

agrobiosystems in which grasslands, hedges and woodlands<br />

occur together. The apparent phenological synchronization<br />

observed in the Dolomites between butterfly cycles and<br />

traditional mowing cycles (Balletto et al. 1988) may be another<br />

example <strong>of</strong> such a process.<br />

Air pollution<br />

Almost every author identifies air pollution as a cause <strong>of</strong><br />

butterfly decline, but this is not supported by detailed evidence<br />

in any case (Thomas 1984b). The effect <strong>of</strong> pollution on insects<br />

is better understood in soil or aquatic species, although it seems<br />

obvious that it may have a real effect upon terrestrial insects. A<br />

dispar, arion, semiargus<br />

arion, teleius, nausithous, hippothoe, semiargus<br />

dispar<br />

argiades, idas, dorylas<br />

28<br />

Total extinct<br />

butterflies<br />

high diversity and population number have been recorded on<br />

road verges in the United Kingdom (Munguira and Thomas<br />

1992), a habitat where pollution caused by car exhausts is<br />

expected to be particularly severe. This probably suggests that,<br />

at least for some species, direct effects <strong>of</strong> pollution are negligible,<br />

whereas indirect effects (for example, <strong>of</strong> acid rain on forests)<br />

may prove to be more important for butterfly populations.<br />

Chemicals (pesticides, herbicides, fertilizers)<br />

A negative effect on insect diversity is <strong>of</strong>ten attributed to any<br />

chemicals sprayed over natural or semi-natural habitats. The<br />

effect <strong>of</strong> pesticides has only been studied on common species,<br />

but it is also evident that no rare or endangered species can<br />

survive on heavily sprayed localities. Some indirect evidence<br />

for this can be found in the literature. In the Padano-venetian<br />

plains, Balletto et al. (1982e) found considerable differences<br />

between butterfly communities in heavily sprayed and<br />

traditionally managed rice fields. Whereas on average five<br />

butterfly species were observed on each <strong>of</strong> the relatively natural<br />

5<br />

15<br />

2<br />

8


habitats which supported Lycaena dispar, only one or two<br />

species could be found in the more heavily sprayed fields and<br />

here L, dispar was totally absent. Other unpublished observations<br />

give further evidence <strong>of</strong> this effect. The conclusion is that<br />

chemicals can reduce butterfly diversity and that they may be<br />

implicated in the disappearance <strong>of</strong> L. dispar from vast areas in<br />

the region.<br />

Erhardt (1985) showed a negative effect <strong>of</strong> grassland<br />

fertilization: whereas six lycaenids among 30 Lepidoptera were<br />

more abundant in an unfertilized meadow than in unfertilized<br />

areas only one lycaenid out <strong>of</strong> two Lepidoptera was more<br />

abundant in fertilized meadows. Butterfly diversity was also<br />

drastically reduced in fertilized compared with unfertilized<br />

meadows. Balletto et al. (1988), however, failed to demonstrate<br />

the same effect on six fertilized and three unfertilized plots in<br />

the Dolomites.<br />

Climatic change<br />

Any relationship between climate and butterfly fluctuations is<br />

also hard to establish. In the study <strong>of</strong> this process, authors have<br />

referred to short-term climatic fluctuations, which are the only<br />

ones we can analyse. Even this needs a database maintained<br />

over several years, and only the British Butterfly Monitoring<br />

Scheme (BMS) is now available for such studies. Pollard<br />

(1988) stated that some climatic parameters such as summer<br />

temperatures are correlated with high butterfly numbers. Climate<br />

can also represent an important factor when fluctuations in<br />

numbers occur in populations previously isolated by other<br />

means such as habitat destruction. The synergetic effect <strong>of</strong> both<br />

factors can certainly endanger populations that are already<br />

declining. For example, a coincidence <strong>of</strong> habitat damage and<br />

unfavourable weather accelerated the extinction <strong>of</strong> Maculinea<br />

arion in Great Britain (Thomas 1989). Long-term climatic<br />

changes normally represent a natural process, but it is now<br />

uncertain if the short-term fluctuations derived from the<br />

greenhouse effect will have any influence on butterfly<br />

populations. This certainly adds another factor probably having<br />

some effect on butterfly abundance or survival, particularly<br />

when considering rare species.<br />

Tourism and urbanisation<br />

Ever-expanding tourist facilities and advancing urbanisation<br />

are obviously among the factors that threaten many butterfly<br />

populations. Nevertheless their effects are far more restricted<br />

geographically than habitat destruction or alteration due to<br />

changes in land management practices. Tourism is particularly<br />

aggressive in some areas, such as the Mediterranean coast, the<br />

Alps or some parts <strong>of</strong> the Pyrenees, where huge areas have<br />

literally been covered by urbanisation or ski courses.<br />

There is no particular lycaenid restricted to coastal areas<br />

around the Mediterranean, but the effect <strong>of</strong> tourism clearly<br />

makes the species' habitats smaller, acting together with other<br />

more extensive impacts. In northwestern Italy Glaucopsyche<br />

melanops (Boisduval) and Satyrium esculi (Hübner) are<br />

29<br />

threatened by tourist resorts which are now spreading inland<br />

from the sea borders (already totally covered in tarmac and<br />

concrete). Some previously abundant populations <strong>of</strong> Lycaena<br />

thesarmon (Esper) have become extinct as a consequence <strong>of</strong> the<br />

spreading urbanisation around Rome.<br />

High mountain habitats are particularly susceptible to the<br />

impact <strong>of</strong> expanding tourist facilities because they host scarce<br />

and ecologically specialised forms. One example is represented<br />

by Vacciniina optilete in the Alps, which is declining in many<br />

parts <strong>of</strong> its former range (Balletto in press). Another example is<br />

represented by Agriades zullichi and Polyommatus golgus,<br />

very rare endemic lycaenids in Sierra Nevada (southern Spain).<br />

Their range is already restricted by a road and a ski resort<br />

development, but they may now disappear from one <strong>of</strong> their<br />

localities if the planned redevelopments really take place.<br />

Collecting and commerce<br />

Again, every review on the causes <strong>of</strong> butterfly decline and<br />

extinction deals with this topic, but appropriate studies on the<br />

effects <strong>of</strong> collectors on butterfly populations remain wanting.<br />

The appeal <strong>of</strong> this topic probably has something to do with<br />

ethics and with the fact that treating some sophisticated and<br />

non-renewable products <strong>of</strong> nature as items <strong>of</strong> commerce is now<br />

unacceptable; neither does it seem correct to kill animals that<br />

other agencies are striving to conserve. Some collecting is still<br />

necessary in areas where our faunistic knowledge is poor.<br />

Sometimes forbidding collection does little if any good for<br />

butterflies (Kudrna 1986). Results obtained from the<br />

enforcement <strong>of</strong> bans on butterfly collection in some cases have<br />

shown this to be an unsuccessful management practice: in<br />

Germany a ban on the collection <strong>of</strong> four butterfly species passed<br />

in 1936 has not prevented the dramatic decline <strong>of</strong> these species<br />

in the course <strong>of</strong> the last 55 years (Kudrna 1989).<br />

In large populations the number <strong>of</strong> butterflies a collector can<br />

take is really negligible, not reaching 10% <strong>of</strong> the total daily<br />

population estimates, while small populations are normally <strong>of</strong><br />

little interest to commercial collectors. To destroy one <strong>of</strong> these<br />

small populations by collection it is necessary to kill almost all<br />

the butterflies seen during the flight period: this represents a<br />

highly time-consuming job with slight rewards for the collector<br />

(Munguira et al. in press).<br />

Legislation to protect species and habitats<br />

In the last 30 years some European countries have passed laws<br />

protecting butterfly species. Heath (1981) reviewed this topic<br />

gathering data from 25 European countries, 13 (52%) <strong>of</strong> which<br />

had some legislation on the matter while only three countries<br />

(12%) included lycaenids among the protected species. In this<br />

review protected lycaenids were Maculinea alcon, M. alcon, M.<br />

teleius, Lycaena dispar, L. helle and Polyommatus bellargus.<br />

Most <strong>of</strong> this legislation has been ineffective because it was based<br />

on the species themselves and paid little attention to habitats.


At the European level, the Council <strong>of</strong> Europe, an international<br />

organisation based in Strasbourg (France) that groups together<br />

almost all European countries, has endeavoured in recent years<br />

to provide collective tools for animal and plant conservation.<br />

This organisation sponsored the publication <strong>of</strong> Heath's<br />

Threatened butterflies in Europe (1981) and Collins and Wells'<br />

Invertebrates in need <strong>of</strong> Special Protection in Europe (1987).<br />

Partially as a result <strong>of</strong> these two reviews some invertebrates<br />

were included in the appendices to the Berne Convention<br />

(Convention on the <strong>Conservation</strong> <strong>of</strong> European Wildlife and<br />

Natural Habitats) in 1988. The Convention has been signed by<br />

18 States and is important because it takes into account the need<br />

for habitat protection together with species protection. Appendix<br />

II <strong>of</strong> the Convention protects 24 lepidopteran species, five <strong>of</strong><br />

which are lycaenids (Lycaena dispar, Maculinea arion, M.<br />

teleius, M. nausithous and Polyommatus golgus).<br />

Another potentially useful international convention yet to<br />

be considered in this framework is the International Convention<br />

for the <strong>Conservation</strong> <strong>of</strong> Wetlands (Ramsar), 1971. Devised to<br />

ensure protection <strong>of</strong> all European important wetlands, Ramsar<br />

legislation is, in principle, an ideal instrument for the<br />

conservation <strong>of</strong> all wet meadows and wetlands including any<br />

threatened lycaenid species. The problem with this convention<br />

is that it was aimed at the conservation <strong>of</strong> birds and has been<br />

opened only in recent times to include other vertebrates<br />

(amphibians, reptiles). The possibility that butterflies may be<br />

also considered in the selection <strong>of</strong> Ramsar sites is unfortunately<br />

nowhere in sight, but insect conservationists should be aware<br />

that such a possibility exists for the future.<br />

Species used as emblems, highlights <strong>of</strong> the<br />

fauna<br />

Some individual species, generally characterised by a<br />

particularly striking appearance, have <strong>of</strong>ten been used to attract<br />

the attention <strong>of</strong> the general public toward conservation issues.<br />

These so-called 'panoramic species' have proved useful to<br />

promote the conservation <strong>of</strong> many animal groups.<br />

Although the importance <strong>of</strong> lycaenids from this point <strong>of</strong><br />

view is not as clear as with papilionids (e.g. Parnasius apollo<br />

(L.), Papilio hospiton Géné) some species have been<br />

instrumental in arousing considerable attention from the public<br />

towards butterfly conservation. The best example for this is<br />

Maculinea arion in Great Britain. During the 1970s the dramatic<br />

decline <strong>of</strong> this species in its few remaining biotopes was closely<br />

followed by many amateur lepidopterists and the topic appeared<br />

in several local and national press releases. A fund was<br />

established to support the preservation <strong>of</strong> the species and<br />

sponsor scientific research on its habitat requirements.<br />

Unfortunately all this interest came too late to save the species<br />

which became extinct in 1979 (Thomas 1989). However, this<br />

process helped in advancing the knowledge <strong>of</strong> the biology <strong>of</strong><br />

the species to such an extent that it was successfully<br />

reintroduced to some sites a few years later. Since then, the<br />

30<br />

peculiarities <strong>of</strong> the biology <strong>of</strong> the 'large blue' and <strong>of</strong> the<br />

Maculinea species in general have fascinated the public and<br />

they are now certainly emblems <strong>of</strong> European efforts to preserve<br />

butterflies.<br />

The reasons for the success <strong>of</strong> Maculinea species in this<br />

respect have something to do with their relatively large size and<br />

beauty among lycaenids, their obligate dependence on Myrmica<br />

ants and the fact that they are very sensitive to subtle changes<br />

in habitat quality (a perfect illustration <strong>of</strong> the reasons for<br />

butterfly decline in Europe). Maculinea species are subjects <strong>of</strong><br />

conservation studies throughout Europe from Spain to Hungary<br />

and are being reintroduced in several countries where they had<br />

previously become extinct. They are also currently included in<br />

every European list <strong>of</strong> endangered Lepidoptera.<br />

Another emblematic lycaenid, from the conservation point<br />

<strong>of</strong> view, is Lycaena dispar, the first butterfly whose extinction<br />

was documented in the United Kingdom as early as 1851. The<br />

habitat <strong>of</strong> this species consists <strong>of</strong> marshes, fens, wet meadows<br />

and oxbow lakes. Drainage <strong>of</strong> such biotopes has been the main<br />

cause <strong>of</strong> its decline throughout its range. It is protected by<br />

national laws in the following seven countries: United Kingdom,<br />

France, Netherlands, Hungary, Germany, Finland and Belgium;<br />

and also by all the Contracting Parties <strong>of</strong> the Berne Convention<br />

(Collins 1987). Apart from the United Kingdom it has also<br />

become extinct in Denmark and endangered in all the rest <strong>of</strong> its<br />

European range. Lycaena dispar is emblematic because it is<br />

sensitive to a very common practice in Europe, wetland drainage,<br />

and because it is conveniently large in size and colourful like the<br />

Maculinea species. It is the only lycaenid included by Kudrna<br />

(1986) in his short list <strong>of</strong> endangered European butterflies.<br />

Species <strong>of</strong> economic importance as pests<br />

There are no real pests among European lycaenids. Only some<br />

very common species can live on Mediterranean legume crops<br />

and could be considered as potential enemies <strong>of</strong> cultivated<br />

plants, but their presence has never caused serious problems.<br />

Lampides boeticus (L.) is occasionally listed among pests <strong>of</strong><br />

peas, and together with Polyommatus icarus and Leptotes<br />

pirithous (L.) can also be a potential pest <strong>of</strong> alfalfa crops<br />

(Martin 1984).<br />

Red Data List <strong>of</strong> European lycaenids<br />

Any Red Data List is subjective and normally tends to be<br />

influenced by the authors' personal experience. Even the Red<br />

Data List concept is controversial and many authors have<br />

expressed opinions against it (Diamond 1988). Some authors<br />

have suggested that a list <strong>of</strong> endangered habitats may be far<br />

more useful than species orientated red data lists (Balletto and<br />

Casale 1991), and Kudrna (1986) has proposed that the latter<br />

should only include species which are in need <strong>of</strong> emergency


action. In our opinion, policies concentrating on habitat<br />

conservation should be given priority over species-centred<br />

schemes although it must be remembered that it is sometimes<br />

easier for policy makers to focus on the protection <strong>of</strong> a species<br />

rather than the more complicated process <strong>of</strong> habitat protection.<br />

In general, we consider that our section on important habitats<br />

for lycaenids is more relevant than the present paragraph on<br />

threatened lycaenids.<br />

In an attempt to make our list as impartial as possible we<br />

have pooled data from five European-level compilations <strong>of</strong><br />

threatened species (Heath 1981; Kudrna 1986; Collins and<br />

Wells 1987; the Berne Convention (Appendix II, 1988) and an<br />

unpublished list drafted by Van der Made during the Wageningen<br />

Symposium on 'Status <strong>of</strong> <strong>Butterflies</strong> in Europe' in 1989). For<br />

every species appearing on a list we have given a score depending<br />

on whether it was listed as endangered (3), vulnerable (2), or<br />

rare (1). Species appearing as indeterminate or other categories<br />

were not considered, and the Berne Convention species<br />

(Appendix II) have all been scored as 3. In this way the highest<br />

possible score for any species is 15 (if 'endangered' on all five<br />

lists) and the lowest is 1.<br />

With a few exceptions that will be dealt with later the results<br />

shown in Table 6 are surprisingly consistent for the high<br />

ranking categories ('endangered' and 'vulnerable' in our list).<br />

All the species listed as 'endangered' have been accorded this<br />

status in at least one <strong>of</strong> the lists referred to above. Species in the<br />

'rare' category are a more mixed group that can be interpreted<br />

as species on which different authors do not agree, or that are<br />

localised but common.<br />

Such a list, however, is far from perfect, mainly because for<br />

the north and centre <strong>of</strong> Europe there is a long tradition in<br />

conservation studies, whereas in the south <strong>of</strong> the continent<br />

comparable research has long been neglected. Polyommatus<br />

coelestinus (Eversmann), for example, is categorised only as<br />

'rare' because it is part <strong>of</strong> a complex which may be abundant in<br />

Anatolia and the Caucasus, but is certainly more than rare in<br />

other parts <strong>of</strong> Europe (e.g. Peloponnisos). There is a similar<br />

situation with Turanana panagea.<br />

Another problem clearly under-ranking some species has to<br />

do with taxonomy (Daugherty et al. 1990). A number <strong>of</strong><br />

Polyommatus <strong>of</strong> the subgenus Agrodiaetus, for instance, have<br />

been recognised as separate species only recently. Accordingly<br />

they were not included in older lists. Another case is Agriades<br />

zullichi which was considered a subspecies <strong>of</strong> the comparatively<br />

common A. glandon (Prunner), a fact that previously obscured<br />

the true endangered status <strong>of</strong> what is now considered a distinct<br />

species. As a consequence, we think that Agriades zullichi,<br />

Polyommatus exuberans and P. humedasae (Toso and Balletto)<br />

should be listed as 'endangered' and Turanana panagea,<br />

Pseudophilotes barbagiae Prins & Poorten, Kretania psylorita,<br />

Polyommatus galloi Balletto & Toso, P. aroaniensis Brown<br />

and P. coelestinus as 'vulnerable'.<br />

31<br />

Priorities in the conservation <strong>of</strong> European<br />

lycaenids<br />

Some conclusions can be drawn from all that has been discussed<br />

in the preceding sections. Most butterfly conservationists agree<br />

that some general considerations are to be kept in mind if any<br />

conservation measure is intended to be effective. Therefore, the<br />

duty <strong>of</strong> ecologists and butterfly conservationists is to make<br />

these available in an understandable way to the people who<br />

make decisions.<br />

The main considerations which emerge from our appraisal<br />

are:<br />

(1) The fact that many lycaenid species are declining is<br />

supported by evidence derived from most European countries.<br />

Habitat destruction and changes in land use are apparently the<br />

two most important factors responsible for species decline. As<br />

a consequence, such alterations should be stopped in the areas<br />

where endangered species have their habitat.<br />

(2) Vulnerable species suffer the same threatening factors<br />

on a more local scale, but any action liable to pose a threat for<br />

them should be carefully considered, and the status <strong>of</strong> the<br />

species over its distributional range taken into account.<br />

(3) Wetlands and wet meadows are among the most<br />

vulnerable habitats in Europe. Accordingly, wetland lycaenids<br />

are unanimously regarded as the most endangered by butterfly<br />

specialists. A European wetland register is urgently needed and<br />

legislation passed in order to protect all European wetlands<br />

known to include populations <strong>of</strong> threatened species. As we have<br />

already said, the most suitable international framework for such<br />

an action is Ramsar which might be persuaded, at a future date,<br />

to incorporate butterfly data into the listing criteria.<br />

(4) Traditional land practices have been compatible for<br />

centuries with the survival <strong>of</strong> species living on those sites and<br />

their continuation may occasionally prove necessary. They<br />

should be maintained wherever they are vital for the survival <strong>of</strong><br />

declining species, and enhanced in those places where they<br />

gave way to more aggressive practices. Nevertheless, this fact<br />

cannot be generalised and some species will certainly do better<br />

in a totally natural habitat.<br />

The policy <strong>of</strong> continuing with traditional practices may<br />

sometimes need subsidies, but may be enough to keep high<br />

standards <strong>of</strong> living in many areas as suggested by the <strong>IUCN</strong><br />

conservation policies, and by several conservation specialists<br />

(Thibodeau and Field 1984). Abandonment <strong>of</strong> former extensive<br />

agricultural areas should be monitored, because this may<br />

eventually lead to changes in habitat as a result <strong>of</strong> natural<br />

succession and the resultant loss <strong>of</strong> many species adapted to<br />

traditional agrobiosystems.<br />

(5) Nature reserves should only be declared for those<br />

habitats and species whose conservation cannot be assured by<br />

current land uses. Special care should be taken to satisfy the<br />

ecological requirements <strong>of</strong> protected species to ensure their<br />

survival. This will mean in many cases that management will<br />

have to oppose the natural succession <strong>of</strong> vegetation. Conflicts<br />

with the requirements <strong>of</strong> other endangered or rare species


Table 6. A Red Data List <strong>of</strong> European <strong>Lycaenidae</strong>, obtained by summing<br />

scores drawn from other previous lists (see text).<br />

ENDANGERED (score 15–8)<br />

Lycaena dispar<br />

Maculinea teleius<br />

M. nausithous<br />

M. alcon<br />

Polyommatus golgus<br />

Maculinea arion<br />

VULNERABLE (score 7–3)<br />

Vacciniina optilete<br />

Lycaena helle<br />

Maculinea rebeli<br />

Cupido lorquinii<br />

Pseudophilotes bavius<br />

Agriades pyrenaicus<br />

Polyommatus humedasae<br />

P. exuberans<br />

Callophrys avis<br />

RARE (score 2–1)<br />

Plebejus pylaon<br />

Polyommatus eroides<br />

P. ainsae<br />

P. galloi<br />

P. violetae<br />

Aricia morronensis<br />

A. eumedon<br />

Pseudophilotes baton<br />

P. barbagiae<br />

Agriades zullichi<br />

Kretania psylorita<br />

Turanana panagea<br />

Cupido carswelli<br />

Iolana iolas<br />

Scolitantides orion<br />

Kretania euripilus<br />

Polyommatus aroaniensis<br />

P. coelestinus<br />

Thesarmonia thetis<br />

(15)<br />

(14)<br />

(14)<br />

(9)<br />

(9)<br />

(8)<br />

(6)<br />

(6)<br />

(4)<br />

(4)<br />

(4)<br />

(3)<br />

(4)<br />

(3)<br />

(3)<br />

(2)<br />

(2)<br />

(2)<br />

(2)<br />

(2)<br />

(2)<br />

(2)<br />

(2)<br />

(2)<br />

(2)<br />

(2)<br />

(2)<br />

(1)<br />

(1)<br />

(1)<br />

(1)<br />

(1)<br />

(1)<br />

(1)<br />

should also be considered in order to make management decisions<br />

which ensure the conservation <strong>of</strong> all target species.<br />

(6) Natural habitats where rare or endangered species live<br />

on climacic plant communities should also be preserved from<br />

any possible alteration by the creation <strong>of</strong> Natural or National<br />

32<br />

Parks. Priority should be given to those places where extremely<br />

localised lycaenids are present, such as the high elevations <strong>of</strong><br />

Sierra Nevada (Spain), or the Idhi Mountain (Crete). Climacic<br />

forests or shrublands must also be preserved when listed species<br />

live on them, but special effort should be made to ensure that the<br />

ecological requirements <strong>of</strong> target species are really those <strong>of</strong> the<br />

climax and do not depart from it in any way, however subtle.<br />

(7) Finally, more research is still needed, even for the most<br />

well known among the European lycaenids. We suggest three<br />

topics <strong>of</strong> special relevance for lycaenid conservation:<br />

• the creation <strong>of</strong> a European database <strong>of</strong> all the areas relevant<br />

for lycaenid conservation with special attention to the needs<br />

<strong>of</strong> endangered or rare species;<br />

• the monitoring and mapping <strong>of</strong> each endangered or<br />

vulnerable species both on a national and European scale;<br />

• the study <strong>of</strong> the ecological requirements <strong>of</strong> every target<br />

species so that sound recommendations can be made to<br />

ensure their long-term survival.<br />

References<br />

BÀLINT, Z. 1991. <strong>Conservation</strong> <strong>of</strong> butterflies in Hungary. In: Kudrna, O.<br />

(Ed.) Schutz der Tagfaltern in Osten Mitteleuropas: Böhmen, Mähren,<br />

Slowakei, Ungarn. Oedippus 3: 5–36.<br />

BÀLINT, Z. and KERTÉSZ, A. 1990. A survey <strong>of</strong> the subgenus Plebejides<br />

(Sauter, 1986) – A preliminary revision. Linn. Belg. 12: 190–224.<br />

BALLETTO, E. (in press). <strong>Butterflies</strong> in Italy: status, problems and prospects.<br />

Proceedings <strong>of</strong> the International Congress 'Future <strong>of</strong> butterflies in Europe',<br />

Wageningen 1989.<br />

BALLETTO, E., BARBERIS, G. and TOSO, G.G. 1982a. Lepidotteri ropaloceri<br />

dei littorali a duna dell-Italia meridionale. Quaderni sulla 'Struttura dell<br />

zoocenosi terrestri' CNR, Roma 3: 153–158.<br />

BALLETTO, E., BARBERIS, G. and TOSO, G.G. 1982b. Aspetti dell'ecologia<br />

dei Lepidotteri ropaloceri nei consorzi erbacei delle Alpi italiane. Quaderni<br />

sulla 'Struttura dell zoocenosi terrestri' CNR, Roma 2(11.2): 11–96.<br />

BALLETTO, E. and CASALE, A. 1991. Mediterranean insect conservation:<br />

the importance <strong>of</strong> pleistocene refugia. In: Collins, N.M. and Thomas, J.A.<br />

(Eds): The conservation <strong>of</strong> insects and their habitats. Academic Press,<br />

London, pp. 121–142.<br />

BALLETTO, E. and KUDRNA, O.1985. Some aspects <strong>of</strong> the conservation <strong>of</strong><br />

butterflies in Italy, with recommendations for a future strategy (Lepidoptera,<br />

Hesperiidae and Papilionoidea). Boll. Soc. Entomol. Ital. 117: 39–59.<br />

BALLETTO, E., LATTES, A. and TOSO, G.G. 1982d. Le comunità di<br />

Lepidotteri ropaloceri come strumento per la classificazione e l'analisi<br />

della qualità degli alti pascoli italiani. Quaderni sulla 'Struttura dell<br />

zoocenosi terrestri' CNR, Roma 2(11.2): 97–139.<br />

BALLETTO, E., LATTES, A. and TOSO, G.G. 1985. An ecological study <strong>of</strong><br />

the Italian rhopalocera. Proc. 3rd Congr. eur. Lepid. Cambridge: 7–14.<br />

BALLETTO, E., LATTES, A., CASSULO, L. and TOSO, G.G. 1988. Studi<br />

sull'ecologia dei Lepidotteri ropaloceri in alcuni ambienti delle Dolomiti.<br />

In: Brandmayr, P. (Ed.) Le Dolomiti: Val di Fiemme – Pale di San Martino.<br />

Zoocenosi e Paesaggio I. Studi trentini di Sci. nat. 64: (suppl.): 87–124.<br />

BALLETTO, E. and TOSO, G.G. 1979. On a new species <strong>of</strong> Agrodiaetus<br />

(<strong>Lycaenidae</strong>) from Southern Italy. Nota lepid. 2: 13–25.<br />

BALLETTO, E., TOSO, G.G. and BARBERIS, G. 1982c. Le comunità di<br />

Lepidotteri ropaloceri nei consori erbacei dell'Appennino. Quaderni sulla<br />

'Struttura dell zoocenosi terrestri' CNR, Roma 2(11.1): 77–143.<br />

BALLETTO, E., TOSO, G.G. and BARBERIS, G. 1982e. Le comunità di<br />

Lepidotteri ropaloceri di alcuni ambienti relitti della Padania. Quaderni<br />

sulla 'Struttura dell zoocenosi terrestri' CNR, Roma 4: 45–67.


BALLETTO, E., TOSO, G.G., BARBERIS, G. and ROSSARO, B. 1977.<br />

Aspetti dell'ecologia dei Lepidoiteri ropaloceri nei consorci erbacei alto<br />

Apenninici. Animalia 4: 277–343.<br />

BALLETTO, E., TOSO, G.G. and LATTES, A. 1989. Studi sulle comunità di<br />

Lepidotteri ropaloceri dell littorale tirrenico. Boll. Museilst. biol. Università<br />

Genova 53: 141–186.<br />

BALLETTO, E., TOSO, G.G. and TROIANO, G. 1981. Aricia cramera<br />

(Escholtz, 1821) in Sardinia. Nota lepid. 4: 81–92.<br />

BIGOT, K. 1952. Biogéographie des lepidoptères de la Provence Occidentale.<br />

Vie Milieu 8(3): 253–264.<br />

BIGOT, K. 1956. Biogéographie des lepidoptères de la Provence Occidentale.<br />

Vie Milieu 8(3): 253–264.<br />

BLAB, J. and KUDRNA, 0.1982. Hilfprogram für Schmetterlinge. Ökologie<br />

und Schutz von Tagfaltern und Widderchen. Kild Verlag; Naturschutz<br />

aktuell 6: 1–135.<br />

CHAPMAN, T.A. 1914. A contribution to the life history <strong>of</strong> Agriades<br />

thersites Cantener. Trans. ent. Soc. Lond. 1914: 285–308.<br />

CLÉU, H. 1950. Les faunes entomologiques méditerranéenes dans le bassin<br />

du Rhône et leurs rapports avec les éléments de la flore. Méms Mus. nation.<br />

Hist. nat. Paris 30: 243–266.<br />

CLÉU, H. 1957. Lépidoptères et biocénoses des génévriers dans les<br />

peuplements du bassin du Rhône. Ann. Soc. ent. Fr. 126: 1–29.<br />

COLLINS, N.M. 1987. Legislation to conserve insects in Europe. Pamphlet<br />

13. Amateur Entomologists' Society, London.<br />

COLLINS, N.M. and WELLS, S.M. 1987. Invertebrates in need <strong>of</strong> special<br />

protection in Europe. Council <strong>of</strong> Europe, Strasbourg.<br />

DAUGHERTY, C.H., CREE, A., HAY, J.M. and THOMPSON, M.B. 1990.<br />

Neglected taxonomy and continuing extinctions <strong>of</strong> tuatara (Sphenodon).<br />

Nature 347: 177–179.<br />

DEMPSTER, J. 1971. Some observations on a population <strong>of</strong> the small copper<br />

butterfly Lycaena phlaeas (Linnaeus) (Lep., <strong>Lycaenidae</strong>). Entomologist's<br />

Gaz. 22: 199–204.<br />

DAVIS, G., FRASER, J. and TYNAN, A. 1958. Population numbers in a<br />

colony <strong>of</strong> Lysandra bellargus Rott. (Lepidoptera). Proc. R. ent. Soc. Lond.<br />

33: 31-36.<br />

DENNIS, R.L.H. 1984. Egg-laying sites <strong>of</strong> the common blue butterfly,<br />

Polyommatus icarus (Rottemburg) (Lepidoptera: <strong>Lycaenidae</strong>): the edge<br />

effect and beyond the edge. Entomologist's Gaz. 35: 85–93.<br />

DIAMOND, J.M. 1988. <strong>Conservation</strong> biology: red books or green lists?<br />

Nature 332: 304-5.<br />

DOUWES, P. 1975. Distribution <strong>of</strong> a population <strong>of</strong> the butterfly Heodes<br />

virgaureae (Lep., <strong>Lycaenidae</strong>). Oikos 26: 332–40.<br />

DUFAY, C. 1961. Faune terrestre et d'eau douce des Pyrénées Orientales. Vie<br />

Milieu 12 (1): suppl.<br />

DUFAY, C. 1965–66. Contribution à la connaissance des peuplements en<br />

Lépidoptères de la Haute Provence. Bull. mens. Soc. linn. Lyon, 159pp.<br />

DUFFEY, E. 1968. Ecological studies on the large copper butterfly Lycaena<br />

dispar Haw. batavus Obth. at Woodwalton Fen National Natural Reserve,<br />

Huntingdonshire. J. appl. Ecol. 5: 69–96.<br />

EHRLICH, P.R. 1984. The structure and dynamics <strong>of</strong> butterfly populations.<br />

In: Vane-Wright, R.I. and Ackery, P.R. (Eds). The biology <strong>of</strong> butterflies.<br />

Academic Press, London, pp. 25–40.<br />

EMMET, A.M. and HEATH, J. (Eds) 1989. The moths and butterflies <strong>of</strong> Great<br />

Britain and Ireland. Vol. 7 (1). Harley Books, Colchester.<br />

ERHARDT, A. 1985. Diurnal Lepidoptera–sensitive indicators <strong>of</strong> change in<br />

the semi-natural grassland. J. appl. Ecol. 22: 849–62.<br />

ERHARDT, A. and THOMAS, J.A. 1991. Lepidoptera as indicators <strong>of</strong> change<br />

in the semi-natural grasslands <strong>of</strong> lowland and upland Europe. In: Collins,<br />

N.M. and Thomas, J.A. (Eds) The conservation <strong>of</strong> insects and their<br />

habitats. Academic Press, London, pp. 213–236.<br />

FIEDLER, K. 1990a. European and North West African <strong>Lycaenidae</strong> and their<br />

associations with ants. J. Res. Lepid. 28: 239–57.<br />

FIEDLER, K. 1990b. Bemerkungen zur Larvalbiologie von Callophrys rubi<br />

L. (Lepidoptera: <strong>Lycaenidae</strong>). Nachr. entomol. Ver. Apollo 11: 121–41.<br />

FIEDLER, K. 1991. Systematic, evolutionary, and ecological implications <strong>of</strong><br />

myrmecophily within the <strong>Lycaenidae</strong> (Insecta: Lepidoptera: Papilionoidea).<br />

Bonner Zool. Monographien 31: 1–210.<br />

33<br />

FIORI, G. 1957. Strymon ilicis Esp. (Lepidoptera: <strong>Lycaenidae</strong>). Boll. Entom.<br />

Bologna 22: 205–56.<br />

FORD, E.B. 1970. <strong>Butterflies</strong>. Collins, London.<br />

FOUASSIN, M. 1961. Lépidoptéres du littoral Beige. Lambillionea 61:<br />

65–71.<br />

GONSETH, Y. 1987. Atlas de distribution des papillons diurnes de Suisse<br />

(Lepidoptera: Rhopalocera). Centre Suisse de Cartographie de la Faune,<br />

Neuchâtel.<br />

HEATH, J. 1981. Threatened Rhopalocera (<strong>Butterflies</strong>) inEurope. Council <strong>of</strong><br />

Europe, Strasbourg.<br />

HEATH, J., POLLARD, E. and THOMAS, J.A. 1984. Atlas <strong>of</strong> <strong>Butterflies</strong> in<br />

Britain and Ireland. Penguin Books, Harmondsworth.<br />

HOEGH-GULDBERG, O. and JARVIS, F.V.L. 1970. Central and North<br />

European Aricia (Lep., <strong>Lycaenidae</strong>): Relationships, heredity, evolution.<br />

Natur. Jutlandica 15: 7–106.<br />

JANMOULLE, E. 1965. Lépidoptères observes dans les dunes du littoral.<br />

Lambillonea 65: 12.<br />

JARVIS, F.V.L. 1958. Biological notes an Aricia agestis (Schiff.) in Britain.<br />

I. Entomologist's Rec. J. Var. 70: 141–148, 169–178.<br />

JARVIS, F.V.L. 1959. Biological notes on Aricia agestis (Schiff.) in Britain.<br />

III. Entomologist's Rec. J. Var. 71: 169–178.<br />

JORDANO, D., FERNÁNDEZ, J. and RODRIGUEZ, J. 1990. The effect <strong>of</strong><br />

seed predation by Tomares ballus (Lep., <strong>Lycaenidae</strong>) on Astragalus<br />

lusitanicus (Fabaceae): determinants <strong>of</strong> differences among patches. Oikos<br />

57: 250–6.<br />

KRALICEK, M. and POVOLNY, D. 1978. Versuch einer Characteristik der<br />

Lepidopterensynusien als primaren Konsumänten in den Vegetationsstufen<br />

der Tschekoslowakei. Vestn. cekosl. Spolecn. Zool. 42: 273–288.<br />

KRATOCHWIL, A. 1989a. Biozonötische Umschichtungen im Grünland<br />

durch Düngung. Sonder. NNA. Ber. 2: 46–58.<br />

KRATOCHWIL, A. 1989b. Community structure <strong>of</strong> flower visiting insects in<br />

different grassland types in Southern Germany. Spixiana 289–302.<br />

KRATOCHWIL, A. 1989c. Erfassung von Blütenbesuchergemeinschaften<br />

verschiedener Rassengesellschaften in Naturschutzgebiet Taubergiessen<br />

(Oberrheinebene). Sonderd. Ges. Ökol., Göttingen 17: 701–711.<br />

KUDRNA, 0.1986. <strong>Butterflies</strong> <strong>of</strong> Europe. Vol. 8. Aspects <strong>of</strong> the conservation<br />

<strong>of</strong> butterflies in Europe. Aula-Verlag, Wiesbaden.<br />

KUDRNA, O. 1989. <strong>Conservation</strong> <strong>of</strong> butterflies in the Federal Republic <strong>of</strong><br />

Germany since 1980. In: Abstracts Int. Congress 'Future <strong>of</strong> butterflies in<br />

Europe: strategies for survival.' Wageningen. Netherlands.<br />

KULFAN, J. and KULFAN, M. 1991. Die Tagfalterfauna der Slowakei und<br />

ihr Schutz unter besonderer Berücksichtigung der Gebirgsökosysteme. In:<br />

Kudrna, O. (Ed.) Schutz der Tagfaltern in Osten Mitteleuropas: Böhmen,<br />

Mähren, Slowakei, Ungarn. Oedippus 3: 75–102.<br />

LEESTMANS.R. 1984. L'écologieet labiogéographie en Europe de Lycaeides<br />

idas (L., 1761). Linn. belg. 9: 370–408.<br />

LEIGHEB, G., RIBONI, E. and CAMERON-CURRY, V. 1990. Kretania<br />

psylorita Freyer (<strong>Lycaenidae</strong>). Discovery <strong>of</strong> a new locality in Crete. Nota<br />

lepid. 13: 242–245.<br />

MALICKY, H. 1961. Über die Ökologie von Lycaeides idas L., insbesondere<br />

über seine symbiose mit Ameisen. Z. Arb. gem. Österr. Ent. 13: 33–49.<br />

MALICKY, H. 1969a. Übersicht über praimaginalstadien, bionomie und<br />

ökologie der mitteleuropaischen <strong>Lycaenidae</strong> (Lepidoptera). Mitt. Ent.<br />

Ges. Basel 19: 25–91.<br />

MALICKY, H. 1969b. Versucht einer Analyse der ökologischen Beziehungen<br />

zwischen Lycaeniden (Lepidoptera) und Formiciden (Hymenoptera).<br />

Tijdschr. Ent. 112: 213–298.<br />

MANINO, Z., LEIGHEB, G., CAMERON-CURRY, P. and CAMERON-<br />

CURRY, V. 1987. Descrizione degli stadi preimarginali di Agrodiaetus<br />

humedasae Toso & Balletto, 1976 (Lepidoptera, <strong>Lycaenidae</strong>). Boll. Mus.<br />

reg. Sci. nat. Torino 5: 97–101.<br />

MARTIN, J. 1981. Similtudes biológicas y diferencias ecológicas entre<br />

Glaucopsyche alexis (Poda) y Glaucopsyche melanops (Boisduval). (Lep.<br />

<strong>Lycaenidae</strong>). Bol. Est. Central Ecol. 10: 59–70.<br />

MARTÍN, J. 1982. La biología de los licénidos españoles (Lep. Rhopalocera).<br />

Miscelánea Commemorativa Fac. Ciencias U. Autón. Madrid: 1003–1020.<br />

MARTÍN, J. 1984. Biología comparadade Lampides boeticus (L.),Syntarucus


pirithous (L.) y Polyommatus icarus (Rot.) (Lep., <strong>Lycaenidae</strong>). Graellsia<br />

40: 163–93.<br />

MIKKOLA, K. and SPITZER, K. 1983. Lepidoptera associated with peatlands<br />

in Central and Northern Europe: a synthesis. Nota lepid. 6: 216–229.<br />

MORTON, A.C. 1985. The population biology <strong>of</strong> an insect with a restricted<br />

distribution: Cupido minimus. PhD thesis, University <strong>of</strong> Southampton.<br />

MUNGUIRA, M.L. 1989. Biología y biogeografía de los licénidos ibéricos en<br />

peligro de extinción (Lepidoptera, <strong>Lycaenidae</strong>). Serv. Publicaciones U.<br />

Autónoma de Madrid, Madrid.<br />

MUNGUIRA, M.L. and MARTIN, J. 1988. Variabilidad morfológica y<br />

biológica de Aricia morronensis (Ribbe), especie endémica de la Península<br />

Ibérica (Lepidoptera: <strong>Lycaenidae</strong>). Ecologia 2: 343–58.<br />

MUNGUIRA, M.L. and MARTIN, J. 1989. Paralelismo en la biología de tres<br />

especies taxonómicamente próximas y ecológicamente diferenciadas del<br />

género Lysandra: L. dorylas, L. nivescens y L. golgus (Lepidoptera,<br />

<strong>Lycaenidae</strong>). Ecologia 3: 331–52.<br />

MUNGUIRA, M.L. and THOMAS, J.A. (1992). Use <strong>of</strong> road verges by<br />

butterfly and burnet populations, and the effect <strong>of</strong> roads on adult dispersal<br />

and mortality. J. appl. Ecol. 29: 316–329.<br />

MUNGUIRA, M.L., THOMAS, J.A., MARTÍN, J. and ELMES, G.W. (in<br />

press). Population size, dispersal and the vulnerability to collectors <strong>of</strong><br />

three endangered species <strong>of</strong> Maculinea butterfly. Biol. Conserv.<br />

MUNGUIRA, M.L., VIEJO, J.L. and MARTÍN, J. 1988. Distribuci6n<br />

geográfica y biología de Eumedonia eumedon (Esper, 1780) en la Peninsula<br />

Ibérica (Lepidoptera: <strong>Lycaenidae</strong>). Shilap Revta. lepid. 16: 217–229.<br />

NEL, J. 1978. Un élevage de Lysandra hispana H.-S. (Lep. <strong>Lycaenidae</strong>).<br />

Alexanor 10: 317–321.<br />

PELLMYR, O. 1983. Plebian courtship revisited: Studies on the femaleproduced<br />

male behaviour-eliciting signals in Lycaeides idas courtship,<br />

(<strong>Lycaenidae</strong>). J. Res. Lepid. 21: 147–57.<br />

POLLARD, E. 1988. Temperature, rainfall and butterfly numbers. J. appl.<br />

Ecol. 25: 819–828.<br />

RAVENSCROFT, N.O.M. 1990. The ecology and conservation <strong>of</strong> the silverstudded<br />

blue butterfly Plebejus argus L. on the sandlings <strong>of</strong> East Anglia,<br />

England. Biol. Conserv. 53: 21–36.<br />

RODRIGUEZ, J. 1991. Las mariposas del Parque Nacional de Doñana.<br />

Biología y ecología de Cyaniris semiargus y Plebejus argus. PhD thesis.<br />

Universidad de Córdoba.<br />

SBN (Schweizerischer Bund fur Naturschutz) 1987. Tagfalter und ihre<br />

Lebensräume. Holliger, Basel.<br />

SCHURIAN, K.G. 1980. Dauerzuchtversucht mit Lysandra hispana (H.S.,<br />

1852) (Lepidoptera, <strong>Lycaenidae</strong>). Alexanor 11: 196–199.<br />

SOURÉS, B. 1974. Relations et facteurs de répartition de différentes espéces<br />

du genre Lysandra (Lepidoptères, <strong>Lycaenidae</strong>). Entomops 33: 25–32.<br />

SWAAY, C.A.M. van 1990. An assessment <strong>of</strong> the changes in butterfly<br />

34<br />

abundance in The Netherlands during the 20th century. Biol. Conserv. 52:<br />

287–302.<br />

TAX, M.H. 1989. Atlas van der nederlandse dagvlinders. Gravenland,<br />

Wageningen.<br />

THIBODEAU, F.R. and FIELD, H.H. 1984. Sustaining Tomorrow. A Strategy<br />

for World <strong>Conservation</strong> and Development. University Press, New England.<br />

THOMAS, CD. 1985. The status and conservation <strong>of</strong> the butterfly Plebejus<br />

argus, L. (Lepidoptera: <strong>Lycaenidae</strong>) in North West Britain. Biol. Conserv.<br />

33: 29–52.<br />

THOMAS, J. A. 1974. Ecological studies <strong>of</strong> hairstreak butterflies. PhD thesis,<br />

University <strong>of</strong> Leicester.<br />

THOMAS, J.A. 1975. Some observations on the early stages <strong>of</strong> the purple<br />

hairstreak butterfly, Quercusia quercus (Linnaeus) (Lep., <strong>Lycaenidae</strong>).<br />

Entomologist's Gaz. 26: 224–226.<br />

THOMAS, J.A. 1980. Why did the large blue become extinct in Britain? Oryx<br />

15: 243–7.<br />

THOMAS, J.A. 1983. The ecology and conservation <strong>of</strong> Lysandra bellargus<br />

(Lepidoptera: <strong>Lycaenidae</strong>) in Britain. J. appl. Ecol. 20: 59–83.<br />

THOMAS, J.A. 1984a. The behaviour and habitat requirements <strong>of</strong> Maculinea<br />

nausithous (the dusky large blue) and M. teleius (the scarce large blue) in<br />

France. Biol. Conserv. 28: 325–47.<br />

THOMAS, J.A. 1984b. The conservation <strong>of</strong> butterflies in temperate countries:<br />

past efforts and lessons for the future. pp. 333–353. In: Vane-Wright, R.I.<br />

and Ackery, P.R. (Eds), The <strong>Biology</strong> <strong>of</strong> <strong>Butterflies</strong>, London.<br />

THOMAS, J.A. 1989. The return <strong>of</strong> the large blue butterfly. British Wildlife<br />

1: 2–13.<br />

THOMAS, J.A., ELMES, G.W., WARDLAW, J.C. and WOYCIECHOWSKY,<br />

M. 1989. Host specificity among Maculinea butterflies in Myrmica ant<br />

nests. Oecologia 79: 452–457.<br />

UHERKOVICH, A. 1972. Adatok a Dravasik nagylepkefaunajanak<br />

(Makrolepidoptera) ismeeretéhez. Vas. megvei Muz. Ert. 5–6: 115–145.<br />

UHERKOVICH, A. 1975. Adatok Buranya nagylepkefaunajanak ismeeretéhez.<br />

iv. A Villanyi-hegiség nappali lepkëi. Janus Pannonius Muz. Evk. 17–18:<br />

33–42.<br />

UHERKOVICH, A. 1976. Adatok a Dél-Dunanuntul nagylepkefaunajanak<br />

(Makrolepidoptera). Folia ent. Hung. 29: 119–137.<br />

VERITY, R. 1943. Le Farfalle diurne d'Italia. Vol. 2. Marzocco, Firenze.<br />

VIEDMA, M.G. 1984. Consideraciones acerca de la conservaación de<br />

especies de insectos. Real Acad. Ciencias, Madrid.<br />

WEIDEMANN, H.-J. 1986. Tagfalter. Band I. Entwicklung, Lebensweise.<br />

Neumann-Neudamm GmbH and Co. KG, Melsungen.<br />

WIKLUND, C. 1977. Observationer över ägglaggning födosöl och vila hos<br />

donzels blavinge, Aricia nicias scandicus Wahlgr. (Lep., <strong>Lycaenidae</strong>).<br />

Entomol. Tidskr. 98: 1–4.


Overview <strong>of</strong> problems in Japan<br />

Toshiya HIROWATARI<br />

Entomological Laboratory, College <strong>of</strong> Agriculture, University <strong>of</strong> Osaka Prefecture, Sakai, Osaka, 591 Japan<br />

In Japan, there are about 240 resident butterfly species which<br />

comprise palaearctic and oriental faunal elements. Fortunately,<br />

none <strong>of</strong> them has been rendered completely extinct but many<br />

local butterfly colonies seem to have been totally eradicated.<br />

The first legislation to protect butterflies as 'Tennen<br />

Kinenbutu' or 'Natural monuments' was promulgated by the<br />

national government in 1932 for Panchala ganesa (<strong>Lycaenidae</strong>,<br />

Arhopalini) in Nara City. Until now, a total <strong>of</strong> 37 species have<br />

been designated as 'Tennen Kinenbutu' by the national and<br />

local governments. However, in some cases, legislation for the<br />

prohibition <strong>of</strong> collecting without any effective measures for<br />

conservation seems to have been ineffective, especially when a<br />

taxon is designated as a protected species rather than as a local<br />

population with a definable habitat or biotype. In fact, the Nara<br />

population <strong>of</strong> P. ganesa seems to have become extinct without<br />

any precise records because collectors lost interest in studying<br />

protected species. Another case is that <strong>of</strong> Shijimia moorei<br />

(Leech) (<strong>Lycaenidae</strong>, Polyommatini). This species had been<br />

known to occur in east Asia in places such as China and Taiwan.<br />

It was not until 1973 that this species was discovered in Kyushu,<br />

Table 1. <strong>Lycaenidae</strong> from Japan listed by Hama et al. (1989).<br />

Species<br />

Artopoeles pryeri Murray<br />

Coreana raphaelis Oberthur<br />

Niphandra fusca Bremer<br />

Shijimiaeoides divinus Fixcen<br />

Tongeia fisheri Eversmann<br />

Lycaeides subsolana Eversmann<br />

Causes <strong>of</strong> decline<br />

Urbanisation<br />

Urbanisation; deforestation<br />

Urbanisation<br />

Road construction<br />

Larch forestation<br />

Habitat degradation<br />

Orchard and golf course construction<br />

Agriculture and spraying<br />

Factory construction<br />

Flood: foodplant extinction<br />

Urbanisation<br />

Succession<br />

Flood control works<br />

35<br />

Japan: its distribution is extremely local, feeding on Lysionotus<br />

pauceflorus (Gesneriaceae) which usually grows on Quercus<br />

trees (Fagaceae) in humid evergreen forest. Just after that, in<br />

1975, this species was designated as 'Tennen Kinenbutu' by the<br />

national government. In this case, the original colonies <strong>of</strong><br />

Kyushu seem to have been conserved. However, there have<br />

been few additional records from other areas (except one from<br />

Nara, Honshu) because collectors do not publish records <strong>of</strong><br />

protected species even if these are caught. It is believed that<br />

some populations <strong>of</strong> S. moorei other than that in Kyushu<br />

become extinct without any definite records <strong>of</strong> this.<br />

Apart from Shijimia moorei, the decline <strong>of</strong> Japanese<br />

butterflies is attributable to alteration in land management<br />

practices and its effect on butterfly habitat. In Japan, most <strong>of</strong> the<br />

victims, such as Shijimiaeoides divinus Fixcen (<strong>Lycaenidae</strong>,<br />

Polyommatini), Coreana raphaelis Oberthur (<strong>Lycaenidae</strong>,<br />

Theclini) and FabriciananerippeC & R. Felder(Nymphalidae),<br />

depend on habitats such as coppice or grassland which have<br />

been maintained by traditional agricultural practices such as<br />

slash and burn, periodical coppicing for fuels and charcoal<br />

Locality<br />

Setagaya, Tokyo<br />

Kawanishi, Yamagata<br />

Kiso, Nagano; Shiga<br />

Kiso, Nagano; Shiga<br />

Kiso, Nagano<br />

Aomori<br />

Aomori<br />

Azumino, Nagano<br />

Komagano, Nagano; Matsumato, Nagano<br />

Matsumato, Nagano<br />

Azumino, Nagano<br />

Minami-azumi, Nagano


production, grass-cutting, animal husbandry, and so on (Sibatani<br />

1989; Ishii 1990).<br />

In Japan, few projects on butterfly conservation are supported<br />

by governmental funding, but awareness is increasing in many<br />

local research groups. Primary interest on the decline <strong>of</strong><br />

butterflies and campaigns for their conservation have mainly<br />

highlighted rather large and beautiful butterflies, i.e. the<br />

'Luehdorfia <strong>Butterflies</strong>', Luehdorfia japonica Leech and L.<br />

puziloi Ehrsch<strong>of</strong>f (Papilionidae) and the 'Great Purple' Sasakia<br />

charonda (Nymphalidae). However, urgent management is<br />

required for some lycaenid species which have survived in<br />

harmony with the traditional land management that is now<br />

being superceded.<br />

In 1989, the Lepidopterological Society <strong>of</strong> Japan (LSJ) (a<br />

group with nearly 1500 members) published the first volume <strong>of</strong><br />

'Decline and <strong>Conservation</strong> <strong>of</strong> butterflies in Japan' (edited by<br />

Hama, Ishii and Sibatani 1989) which was the first systematic<br />

approach to the problem <strong>of</strong> butterfly conservation in Japan. Six<br />

species <strong>of</strong> <strong>Lycaenidae</strong> were included in that book (Table 1).<br />

Soon after publication, the LSJ held its first seminar on<br />

'<strong>Conservation</strong> <strong>of</strong> <strong>Butterflies</strong>' in June 1990. We found that there<br />

36<br />

are many more local extinctions <strong>of</strong> butterflies than we expected,<br />

and fewer conservation projects supported by adequate budgets.<br />

Attempts to find adequate funds for protecting butterflies or<br />

their habitats and for monitoring and overseeing activities<br />

continue. In early 1990, a group <strong>of</strong> the LSJ was granted funds<br />

<strong>of</strong> JPY 3,450,000 from the Nippon Life Insurance Foundation<br />

for monitoring the latest states <strong>of</strong> distribution and changes in<br />

population size <strong>of</strong> Japanese butterflies as environmental<br />

indicators for quality <strong>of</strong> human life.<br />

References<br />

HAMA, E., ISHII, M. and SIBATANI, A. (Eds) 1989. Decline and conservation<br />

<strong>of</strong> butterflies in Japan. I. Lepidopterological Society <strong>of</strong> Japan, Osaka.<br />

ISHII, M. 1990. What has led to the butterflies declining? Shizen Hogo (The<br />

conservation <strong>of</strong> Nature) 337: 14–15. (In Japanese).<br />

SIBATANI, A. 1989. Decline and conservation <strong>of</strong> butterflies in Japan. In:<br />

Hama, E., Ishii, M. and Sibatani, A. (Eds) Decline and <strong>Conservation</strong> <strong>of</strong><br />

<strong>Butterflies</strong> in Japan. I Lepidopterological Society <strong>of</strong> Japan, Osaka, pp.<br />

16–22.


<strong>Conservation</strong> <strong>of</strong> North American lycaenids – an overview<br />

J. HALL CUSHMAN and Dennis D. MURPHY<br />

Center for <strong>Conservation</strong> <strong>Biology</strong>, Department <strong>of</strong> Biological Sciences, Stanford University, Stanford, California 94305-5020,<br />

U.S.A.<br />

Introduction<br />

Two distinct patterns emerge when one examines the United<br />

States' Endangered Species List. First, although butterflies<br />

probably constitute less than 1 % <strong>of</strong> global insect species richness,<br />

they are disproportionately represented on the list: 53% (14 <strong>of</strong><br />

26) <strong>of</strong> the insects currently afforded federal protection are<br />

butterflies. Second, members <strong>of</strong> the <strong>Lycaenidae</strong> (including the<br />

blues, coppers, hairstreaks and metalmarks) are<br />

disproportionately represented: the family comprises only 21%<br />

<strong>of</strong> the species-level butterfly fauna <strong>of</strong> North America (Scott<br />

1986), but constitutes 50% <strong>of</strong> the listed butterfly taxa: see<br />

Table 1, (Federal Register 1991a).<br />

This over-representation <strong>of</strong> lycaenids also extends to the list<br />

<strong>of</strong> candidate species awaiting protection, including some taxa<br />

at risk <strong>of</strong> imminent extinction (see Table 2). For the U.S. as a<br />

whole, lycaenids comprise 37% <strong>of</strong> all butterflies that are<br />

candidates for listing as endangered species, and this number<br />

rises to 49% if skippers (a group currently treated as a superfamily<br />

distinct from the 'true butterflies') are excluded (Federal Register<br />

1991b). In California, lycaenids are substantially overrepresented<br />

among those taxa listed as candidates for federal<br />

protection or known to be at particular risk, making up 10 <strong>of</strong><br />

those 20 taxa, whereas the state contains only a third <strong>of</strong> the<br />

nation's lycaenid fauna (Murphy 1987a). In addition, the<br />

candidates list is dominated by taxa from California and/or<br />

Nevada, making up 71% <strong>of</strong> the candidates (20 <strong>of</strong> 28; Table 2).<br />

They include a subspecies <strong>of</strong> Plebejus saepiolus (Boisduval)<br />

(soon to be described) that is probably already extinct. Another<br />

undescribed subspecies <strong>of</strong> hairstreak, Incisalia mossii (Edwards),<br />

has been pushed towards extinction by overzealous collectors<br />

who have removed hostplants and larvae. Also included is a<br />

copper, Lycaena hermes (Edwards), that may be differentiated<br />

sufficiently from all known relatives to warrant its recognition<br />

as a monotypic genus. It has been extirpated in significant<br />

portions <strong>of</strong> its historical range on both sides <strong>of</strong> the California-<br />

Mexico border.<br />

There are two opposing ways <strong>of</strong> viewing the lycaenid<br />

dominance <strong>of</strong> endangered and threatened species, and the list <strong>of</strong><br />

candidates for this status. The first interpretation <strong>of</strong> this pattern<br />

37<br />

is that it results from the biased study <strong>of</strong> U.S. butterfly families.<br />

It may be that non-lycaenid butterflies are just as endangered as<br />

lycaenids, but the latter have received more attention from<br />

biologists, and thus more is known about their imperilled state.<br />

While it is difficult to evaluate this contention, there are no<br />

obvious reasons why lycaenids should have received more<br />

attention than other families, given that they are small in size<br />

and not nearly as showy. We favour a second interpretation,<br />

which is that the patterns reflect ecological differences among<br />

the butterfly families.<br />

Here, we first provide a brief overview <strong>of</strong> the taxonomic and<br />

geographic distributions <strong>of</strong> North American lycaenids. We then<br />

discuss five interrelated characteristics <strong>of</strong> lycaenids that we<br />

suspect are responsible for, or at least contribute to, the group's<br />

extreme susceptibility to endangerment and extinction. We<br />

conclude by summarising the ongoing efforts to conserve North<br />

American lycaenids. Throughout the chapter, we focus primarily<br />

on lycaenids and conservation programmes in the U.S. In part<br />

this is a reflection <strong>of</strong> our experience with lycaenids in this<br />

region. However, our restricted emphasis also occurs because<br />

insect conservation in Canada and Mexico is far less developed<br />

(see review by Opler 1991). Particularly in Mexico, a great<br />

number <strong>of</strong> insect taxa are at risk, but specific details are lacking.<br />

Taxonomic and geographic distributions<br />

Lycaenids are somewhat under-represented in North America<br />

and many <strong>of</strong> them just barely enter the United States from<br />

Mexico. The subfamily Riodininae (the metalmarks) is especially<br />

under-represented when compared to the equatorial latitudes in<br />

the New World, with just 20 species (14%) in North America.<br />

By comparison, riodinines make up about two-thirds <strong>of</strong> the<br />

local lycaenid fauna in equatorial lowland communities <strong>of</strong><br />

South America. Indeed, the genera that include all but three <strong>of</strong><br />

the North American species – Apodemia Felder & Felder,<br />

Calephelis Grote & Robinson, and Emesis F. – reach much<br />

greater species richness to the south <strong>of</strong> the United States.<br />

The subfamily Lycaeninae (the harvesters, hairstreaks,<br />

coppers, and blues) is represented by more than 120 species,


Table 1. Lycaenid taxa that are listed as endangered by the U.S. government as <strong>of</strong> 15 July 1991 (Federal Register 1991a). CA = California stale.<br />

Species<br />

Apodemia mormo langei<br />

(Lange's Metalmark)<br />

Glaucopsyche lygdamus palosverdesensis<br />

(Palos Verdes Blue)<br />

Euphilotes baltoides allyni<br />

(El Segundo Blue)<br />

Euphilotes enoptes smithi<br />

(Smith's Blue)<br />

Icaricia (=Plebejus) icarioides missionensis<br />

(Mission Blue)<br />

Lycaeides argyrognomon (=idas) lotis<br />

(Lotis Blue)<br />

Incisalia (=Callophrys) mossii bayensis<br />

(San Bruno Elfin)<br />

Range<br />

including some endemic species groups. Feniseca Grote is an<br />

endemic, monotypic genus <strong>of</strong> the small tribe <strong>of</strong> harvesters<br />

(Miletini) whose predaceous larvae feed on homopterans.<br />

Among the hairstreaks (Theclini) are a number <strong>of</strong> largely<br />

neotropical genera that are represented by one or just several<br />

species, including Eumaeus Hübner, Atlides Hübner,<br />

Chlorostrymon Clench, Tmolus Hübner, Calycopis Scudder,<br />

Cyanophrys Clench, Strymon Hübner, Erora Scudder and<br />

others. The genus Callophrys Billberg (sensu stricto) is holarctic.<br />

Genera that are largely restricted to North America (and mainly<br />

distributed in the west) include the species-rich Satyrium Scudder,<br />

Mitoura Scudder, and Incisalia Scudder (lumped with Callophrys<br />

by many), and the striking monotypic Sandia Clench & Ehrlich.<br />

Representatives <strong>of</strong> the coppers (Lycaenini) are holarctic in<br />

distribution, but many representatives are endemic to North<br />

America. Two species are related to Old World species, Lycaena<br />

cupreus (Edwards) and L. phlaeas (L.), the latter being found<br />

also in Asia and Africa. Three other species are represented in<br />

eastern North America. The remaining eight species, with the<br />

exception <strong>of</strong> the aforementioned highly restricted Lycaena<br />

hermes, are distributed across the intermountain west.<br />

Like the hairstreaks, the blues (Polyommatini) include a<br />

number <strong>of</strong> genera that are more diverse in the neotropics, just<br />

barely reaching into the U.S. (i.e. Hemiargus Hübner and<br />

Leptotes Scudder), and genera that share a similar distribution,<br />

but are also represented in Africa (i.e. Brephidium Scudder and<br />

Zizula Chapman). Celastrina Tutt and Everes Hübner are<br />

holarctic in distribution as functionally are Glaucopsyche<br />

Scudder and Plebejus Kluk (includingAgriades Hübner, Icaricia<br />

Nabokov, Lycaeides Hübner, and Plebulina Nabokov), genera<br />

that show high affinity with many European genera (i.e.<br />

Maculinea van Ecke, Aricia Reichenbach, Polyommatus Kluk,<br />

Plebicula Higgins, Lysandra Hemming, and Agrodiaetus<br />

Hübner, among others). North American Philotiella Mattoni,<br />

the monotypic Philotes Scudder, and the highly subspeciated<br />

Euphilotes Mattoni are all related to Asian genera.<br />

CA<br />

CA<br />

CA<br />

CA<br />

CA<br />

CA<br />

CA<br />

38<br />

Food Plant<br />

Eriogonum nudum<br />

Astragalus trichopodus var. lonchus<br />

Eriogonum parvifolium, E. cinereum, E. fasciculatum<br />

Eriogonum latifolium, E. parvifolium<br />

Lupinus albifrons, L. formosus, L. variicolor<br />

Lotus formosissimus<br />

Sedum spathulifolium<br />

Regional numbers indicate that lycaenids make up a higher<br />

proportion <strong>of</strong> butterfly species richness in the far western<br />

portions <strong>of</strong> the continent. For example, they comprise 29 <strong>of</strong> the<br />

162 (18%) butterfly and skipper species in Georgia (Harris<br />

1972) and 46 <strong>of</strong> the 240 (19%) species in Colorado (Brown<br />

1957, corrected for recent taxonomic changes). For southern<br />

California, Emmel and Emmel (1973) list 167 species <strong>of</strong> which<br />

54 (32%) are lycaenids. On a narrower geographic scale,<br />

lycaenids make up 37 <strong>of</strong> the 122 (30%) butterfly species from<br />

the San Francisco Bay area (Tilden 1965) and 30 <strong>of</strong> the 91<br />

(33%) species from Orange County in southern California<br />

(Orsak 1978).<br />

Perhaps not surprisingly, it is in western North America –<br />

with its diverse topography, elevation, and vegetation types –<br />

where differentiation at or below the species level is greatest<br />

(see Scott 1986). Additionally, it is in this same region, which<br />

harbours the most diverse temperate-zone lycaenid genera at<br />

the species level (i.e. Mitoura, Callophrys, Incisalia, Lycaena,<br />

Plebejus, and Euphilotes), that geographically restricted taxa<br />

are receiving the most attention from conservation biologists.<br />

Lycaenid characteristics and<br />

susceptibility to endangerment<br />

Subspecific differentiation<br />

As can be seen in Tables 1 and 2, 100% <strong>of</strong> the lycaenids<br />

currently protected under the U.S. Endangered Species Act<br />

(ESA), and 88% <strong>of</strong> the candidates for this status, are not full<br />

species. In a striking show <strong>of</strong> appreciation for the process <strong>of</strong><br />

evolution, the U.S. Congress included both full species and<br />

subspecies as protectable taxonomic units (for vertebrates,<br />

even distinct populations at risk <strong>of</strong> endangerment or extinction<br />

are protected). Indeed, the most publicised extinction <strong>of</strong> a North


Table 2. Lycacnid taxa that are on the list <strong>of</strong> candidates for endangered status with the U.S. government.<br />

Data are current as <strong>of</strong> 21 November 1991 (Federal Register 1991b). Abbreviations <strong>of</strong> U.S. states are as follows: California (CA), Florida (FL), Illinois (IL), Indiana<br />

(IN), Maine (ME), Massachusetts (MA), Michigan (MI), New Hampshire (NH), Nevada (NV), New York (NY), Ohio (OH), Oregon (OR), Pennsylvania (PA), and<br />

Wisconsin (WI). Status categories are as follows: 1 = sufficient information to support a proposal for listing as endangered or threatened; 2 = current information<br />

suggests that taxon is 'possibly appropriate' for listing as endangered or threatened; 3A = best available information indicates that the taxon is extinct; 3C = current<br />

information indicates that the taxon is more abundant or widespread than previously thought; S = a taxon known to have stable numbers, U = additional information<br />

is required to determine the taxon's current abundance trend, D = taxon with declining numbers and/or which is being subjected to increasing threats.<br />

Species<br />

Eumaeus atala florida<br />

(Florida Atala)<br />

Euphiloles battoides spp.<br />

(Baking Powder Flat Blue)<br />

Euphilotes enoptes spp.<br />

(Dark Blue)<br />

Euphilotes rita spp.<br />

(Sand Mountain Blue)<br />

Euphilotes rita mattoni<br />

(Mattoni Blue)<br />

Hemiargus thomasi bethunebakeri<br />

(Miami Blue)<br />

Icaricia (=Plebejus) icarioides spp.<br />

(Point Reyes Blue)<br />

Icaricia (=Plebejus) icarioides spp.<br />

(White Mountains Icarioides Blue)<br />

Icaricia (=Plebejus) icarioides spp.<br />

(Spring Mountains Icarioides Blue)<br />

Icaricia (=Plebejus) icarioides fenderi<br />

(Fender's Blue)<br />

Icaricia (=Plebejus) icarioides moroensis<br />

(Morro Bay Blue)<br />

Icaricia (=Plebejus) icarioides pheres<br />

(Pheres Blue)<br />

Incisalia lanoraieensis<br />

(Spruce-Bog Elfin)<br />

Incisalia mossii ssp. ('hikupa')<br />

(San Gabriel Mountains Blue)<br />

Incisalia mossii ssp.<br />

(Marin Elfin)<br />

Lycaeides melissa samuelis<br />

(Karner Blue)<br />

Lycaeides dorcas claytoni<br />

(Clayton's Copper)<br />

Lycaena hermes<br />

(Hermes Copper)<br />

Lycaena rubidus spp.<br />

(White Mountains Copper)<br />

Mitoura gryneus sweadneri<br />

(Sweadner's Olive Hairstreak)<br />

Mitoura thornei<br />

(Thome's Hairstreak)<br />

Philotiella speciosa bohartorum<br />

(Bohart's Blue)<br />

Plebulina emigdionis<br />

(San Emigdio Blue)<br />

Plebejus saepiolus spp.<br />

(San Gabriel Mountains Blue)<br />

Status<br />

2,S<br />

2,U<br />

2,U<br />

2,U<br />

2,U<br />

3C,U<br />

2,U<br />

2,U<br />

NV<br />

2,U<br />

NV<br />

2,U<br />

2,U<br />

3A<br />

3C<br />

2,U<br />

2,U<br />

1,D<br />

2,S<br />

2,U<br />

2,U<br />

2,U<br />

2,U<br />

2,U<br />

2,U<br />

2,U<br />

39<br />

Range<br />

FL<br />

NV<br />

NV<br />

NV<br />

NV<br />

FL<br />

CA<br />

CA,<br />

CA,<br />

OR<br />

CA<br />

CA<br />

ME,NY,NH,<br />

Canada<br />

CA<br />

CA<br />

IN,MI,NH,NY,OH<br />

MA,IL,WI,PA.<br />

ME<br />

CA,<br />

Mexico<br />

CA.NV<br />

FL<br />

CA<br />

CA<br />

CA<br />

CA<br />

Food plant<br />

Zamia pumila<br />

Eriogonum sp.<br />

Eriogonum sp.<br />

Eriogonum sp.<br />

Eriogonum sp.<br />

Eriogonum sp.<br />

Lupinus sp.<br />

Lupinus sp.<br />

Lupinus sp.<br />

Lupinus sp.<br />

Lupinus chamissonis<br />

Lupinus sp.<br />

Picea mariana<br />

Sedum sp.<br />

Sedum sp.<br />

Lupinus perennis<br />

Potentilla sp.<br />

Rhamnus crocea<br />

Rumex sp.<br />

Juniperus silicicola<br />

Cupressus forbesii<br />

Chorizanthe membranacea<br />

Atriplex canescens<br />

Lupinus sp.<br />

Continued…


Table 2 (cont). Lycaenid taxa that are on the list <strong>of</strong> candidates for endangered status with the U.S. government.<br />

Species<br />

Plebejus saepiolus spp.<br />

(While Mountains Saepiolus Blue)<br />

Plebejus shasta charlestonensis<br />

(Spring Mountains Blue)<br />

Satyrium auretorum fumosum<br />

(Santa Monica Mountain Hairstreak)<br />

Slrymon acis bartrami<br />

(Bartram's Hairstreak)<br />

Status<br />

2,U<br />

NV<br />

2,D<br />

American butterfly was a putative full species, the Xerces Blue<br />

(Glaucopsychexerces (Boisduval)), a lycaenid that was probably<br />

a well-differentiated subspecies <strong>of</strong> the still widely distributed<br />

and highly variable Glaucopsyche lygdamus (Doubleday).<br />

Lycaenids appear to have a high degree <strong>of</strong> subspecific<br />

variation and these differentiated populations (many formally<br />

recognized as subspecies) are particularly susceptible to<br />

endangerment and extinction. Analysis <strong>of</strong> Scott's (1986)<br />

taxonomically conservative treatment provides empirical<br />

support for this view. According to his classifications, the ratio<br />

<strong>of</strong> subspecies to species in North America is 2.4 times greater<br />

for lycaenids than for non-lycaenid butterflies (187/142 and<br />

303/537 respectively). In addition, this ratio for those lycaenids<br />

largely restricted to western North America (i.e. Mitoura,<br />

Callophrys [including Incisalia], Lycaena, Plebejus,<br />

Glaucopsyche and Euphilotes) is 4.7 times higher than that for<br />

non-lycaenid butterflies (138/52).<br />

Dispersal abilities<br />

Dispersal is essential for the persistence <strong>of</strong> isolated populations.<br />

First, input <strong>of</strong> individuals from neighbouring areas can bolster<br />

populations whose numbers are dwindling, thereby preventing<br />

their extinction (the 'rescue effect' sensu Brown and Kodric-<br />

Brown 1977). Second, dispersal can provide an influx <strong>of</strong><br />

genetically different individuals into a population, thereby<br />

increasing genetic diversity and presumably resulting in greater<br />

fitness and population viability (see review by Vrijenhoek<br />

1985). Thus, taxa with limited dispersal abilities should be far<br />

more susceptible to local extinction events than taxa with welldeveloped<br />

dispersal abilities.<br />

Although there have been numerous mark-recapture studies<br />

<strong>of</strong> North American butterflies (e.g. Ehrlich 1965, Arnold 1983,<br />

Murphy et al. 1986, Reid and Murphy 1986), few have focused<br />

on lycaenids. Despite the fact that many studies tend to<br />

underestimate mean dispersal distances, research to date<br />

indicates that lycaenids tend to move short distances between<br />

captures. For example, studies <strong>of</strong> the endangered Mission Blue<br />

(Plebejus icarioides missionensis Hovanitz) indicate that most<br />

adult movements are highly restricted: the majority <strong>of</strong> captures<br />

were in the immediate vicinity <strong>of</strong> larval hostplants and nectar<br />

resources (see Arnold 1983, Reid and Murphy 1986). However,<br />

studies have also shown a limited number <strong>of</strong> movements in the<br />

2,U<br />

2,U<br />

40<br />

Range<br />

CA,<br />

NV<br />

CA<br />

FL<br />

Food plant<br />

Lupinus sp.<br />

Trifolium, Astragalus,<br />

Lupinus<br />

Quercus sp.<br />

Croton linearis<br />

order <strong>of</strong> hundreds <strong>of</strong> metres, as well as a single dispersal event<br />

between habitat patches (and distinct demographic units) <strong>of</strong><br />

nearly 2 kilometres. Nevertheless, the tendency for lycaenids to<br />

be comparatively sedentary should result in less frequent<br />

recolonisation and rescue events as well as reduced gene flow<br />

between populations, leading to greater interpopulation<br />

differentiation.<br />

Host specificity and successional stages<br />

All lycaenids currently protected in the U.S. are specific to one<br />

or just several related hostplant species (Tables 1 and 2). Many<br />

<strong>of</strong> these plant species, particularly the commonly used<br />

Eriogonum and Lupinus, are found largely in early successional<br />

communities that are temporary and unpredictable. <strong>Butterflies</strong><br />

that specialise on such plants must track an ephemeral resource<br />

base that itself may be dependent on unpredictable and perhaps<br />

infrequent ecosystem disturbances. For such species, suitable<br />

habitat can be a limited, ever-shifting fraction <strong>of</strong> a greater<br />

landscape mosaic. As a result, local extinction events are both<br />

frequent and inevitable.<br />

The endangered Karner Blue (Lycaeides melissa samuelis<br />

Nabokov) is a prime example <strong>of</strong> the costs associated with such<br />

specialisation. Larvae <strong>of</strong> this subspecies feed exclusively on a<br />

lupine (Lupinus perennis) that is an early successional species<br />

restricted to pine-barren habitats (Zaremba 1991). The existence<br />

<strong>of</strong> these habitats is highly dependent on the occurrence <strong>of</strong><br />

intermittent fires. However, in New York State, fire suppression<br />

and habitat loss have significantly reduced the size and number<br />

<strong>of</strong> this butterfly's patchily distributed populations.<br />

Even when appropriate larval hostplants and successional<br />

stages are available, conditions may still be insufficient to<br />

sustain lycaenid populations. For example, the federally listed<br />

San Bruno Elfin (Incisalia mossiibayensis (Brown)) is not only<br />

restricted to a few rocky outcrops that support its narrowly<br />

distributed hostplant (Sedum spathuliifolium), but the butterfly<br />

apparently can only complete its life cycle on individual plants<br />

growing under highly exacting and uncommon topoclimatic<br />

environments (Weiss and Murphy 1990).<br />

Sedentary behaviour combined with high levels <strong>of</strong> sitespecific<br />

hostplant adaptation are likely to place many genetically<br />

distinct lycaenids at great risk <strong>of</strong> local and regional extinction.<br />

While no studies document the existence <strong>of</strong> genetically based


host races in North American lycaenids, recent work details a<br />

genetic basis for larval hostplant preferences by adult butterflies<br />

in the rather sedentary nymphalid genus Euphydryas. Singer<br />

and colleagues have found significant differences in patterns <strong>of</strong><br />

oviposition preference and tolerances for hostplants among<br />

phenotypically and geographically distinct populations,<br />

suggesting distinct adaptation at the population level (see<br />

Singer 1971; White and Singer 1974; Rausher 1982). More<br />

recently, Singer and colleagues have documented the existence<br />

<strong>of</strong> genetically based (i.e. heritable) differences between adjacent<br />

populations and within polyphagous populations (Singer 1983;<br />

Singer et al. 1988).<br />

Species in the lycaenid genus Euphilotes exhibit similar<br />

patterns in the use <strong>of</strong> larval hostplants, thereby suggesting the<br />

possibility <strong>of</strong> genetic differentiation among populations for<br />

hostplant tolerance (see Pratt and Ballmer 1986). In southern<br />

California, Euphilotes enoptes (Behr) has been found to feed on<br />

five species <strong>of</strong> Eriogonum; the butterfly is monophagous at<br />

some locations, while it is polyphagous in others, with clear<br />

host preferences.<br />

Association with ants<br />

Roughly half <strong>of</strong> the lycaenid species world-wide associate with<br />

ants, and their larvae possess numerous distinctive structures<br />

that facilitate these interactions (Downey 1962; Atsatt 1981;<br />

Pierce 1987). Although a few <strong>of</strong> these associations are<br />

antagonistic, with butterfly larvae preying on ant brood (Cottrell<br />

1984), the majority appear to be mutualistic (Pierce 1987).<br />

Lycaenids vary greatly in terms <strong>of</strong> their degree <strong>of</strong> dependence<br />

on ant associates and their degree <strong>of</strong> specificity for particular<br />

ant species.<br />

We propose that butterfly species which associate with ants,<br />

and particularly those species with strong dependence on them,<br />

are far more sensitive to environmental changes and thus more<br />

prone to endangerment and extinction, than species that are not<br />

tended by ants. While this hypothesis remains untested, it seems<br />

probable because <strong>of</strong> two factors. First, such species<br />

simultaneously require the right food plant and the presence <strong>of</strong><br />

particular ant species – a combination that occurs infrequently.<br />

These dual requirements <strong>of</strong> tended species should result in<br />

spatial distributions that are patchier than those for untended<br />

species. The degree <strong>of</strong> patchiness should increase as dependence<br />

and/or the species specificity <strong>of</strong> lycaenids increase. Second, we<br />

suspect that selection will favour reduced dispersal by<br />

myrmecophilous lycaenids, because <strong>of</strong> the difficulty associated<br />

with locating patches that contain the appropriate combination<br />

<strong>of</strong> food plants and ants. Thus, in addition to occurring as<br />

isolated populations <strong>of</strong> variable sizes, ant-tended species may<br />

express genetic traits associated with reduced outcrossing.<br />

At this time, we cannot evaluate whether North American<br />

lycaenids that associate with ants are more vulnerable to<br />

endangerment or extinction than those without such<br />

dependencies. This is because there have been almost no<br />

studies <strong>of</strong> the ant associations <strong>of</strong> endangered lycaenid taxa.<br />

However, Downey (1962) observed that the Mission Blue was<br />

41<br />

tended by the ant Formica lasioides, and suggested (but did not<br />

demonstrate) that ants may protect caterpillars from natural<br />

enemies and even transport them to their food plants. D.A.<br />

Savignano has studied the ant associations <strong>of</strong> the Karner Blue,<br />

but this work has yet to be published.<br />

Numerous studies <strong>of</strong> non-endangered taxa in North America<br />

suggest that ants could be an important factor in the persistence<br />

<strong>of</strong> lycaenid populations. For example, parasitism levels <strong>of</strong><br />

Glaucopsyche lygdamus oro (Scudder) (the Rocky Mountain<br />

subspecific relative <strong>of</strong> the federally listed Palos Verdes Blue, G.<br />

lygdamus palosverdesensis Perkins and Emmel) were 45–84%<br />

lower for ant-tended larvae than for untended larvae (Pierce and<br />

Mead 1981; Pierce and Easteal 1986). In Michigan, the work <strong>of</strong><br />

Webster and Nielson (1984) also suggested that ant associates<br />

were beneficial for the Scrub-oak or Edward's Hairstreak,<br />

Satyrium edwardsii (Grote & Robinson). Clearly, we need to<br />

know much more about the ant associations <strong>of</strong> endangered<br />

lycaenids, as these interactions will be important considerations<br />

in management plans.<br />

Although not from North America, the Large Blue<br />

(Maculinea arion (L.)) provides an important, sobering, example<br />

<strong>of</strong> the <strong>of</strong>ten dire consequences associated with a dependence on<br />

ants (see Thomas 1980; Cottrell 1984; New 1991). Despite<br />

considerable efforts to prevent its loss, in 1979 the Large Blue<br />

became extinct in its native Britain. While many factors<br />

undoubtedly contributed to this demise, the most prominent<br />

appears to have been the species' extreme dependence on ants.<br />

During early instars, M. arion larvae fed on wild thyme (Thymus<br />

drucei praecox) and, at the fourth instar, were carried by<br />

Myrmica ants into their nests, where the lycaenids fed on ant<br />

brood. The level <strong>of</strong> grazing in the blue's grassland habitats was<br />

progressively reduced from around 1950, largely due to changing<br />

agricultural practices and attempts to protect habitat <strong>of</strong> this<br />

endangered species. However, due to unforeseen complexities<br />

<strong>of</strong> the system, these altered grazing regimes had drastic effects<br />

on the lycaenid populations. The primary ant-species host (M.<br />

sabuleti) could persist only in fields that were closely cropped<br />

by livestock. Thus, even slight reductions in grazing allowed M.<br />

scabrinodis, a low-quality host, to exclude M. sabuleti from the<br />

area, thereby leading to the butterfly's subsequent demise.<br />

<strong>Conservation</strong> planning in North America<br />

Lycaenids have played a central role in the development <strong>of</strong><br />

environmental interests over land-use policy. Although much<br />

less publicised than the Large Blue another lycaenid provides<br />

a further example <strong>of</strong> the kind <strong>of</strong> conservation efforts that are<br />

required to protect endangered butterflies.<br />

The Mission Blue was conferred protection under the ESA<br />

in 1976, when the U.S. Fish and Wildlife Service formally<br />

recognized that encroaching urbanisation had virtually encircled<br />

the known distribution <strong>of</strong> this subspecies. More than half <strong>of</strong> the<br />

grassland habitat <strong>of</strong> the largest remaining known population on<br />

California's San Bruno Mountain had been lost during the 50<br />

years preceding the listing. Furthermore, half <strong>of</strong> this remaining<br />

habitat (a quarter <strong>of</strong> the total) had been overtaken by invasive


shrub and tree species. In 1978, developers and local government<br />

became aware that pending development would be prohibited<br />

by Section 9 <strong>of</strong> the ESA.<br />

Given that the Mission Blue occurred primarily on private<br />

land and the ESA only <strong>of</strong>fered remedies for taxa on public<br />

lands, there was a pressing need for an innovative plan that<br />

would balance biological and economic concerns. Such an<br />

approach was engineered by a committee composed <strong>of</strong><br />

developers, environmentalists, government <strong>of</strong>ficials, and<br />

biological consultants. Using size estimates <strong>of</strong> the butterfly<br />

populations, distributional records for three lupine (Lupinus)<br />

larval hostplants, and information on the butterfly's natural<br />

history, the committee designed the first 'Habitat <strong>Conservation</strong><br />

Plan' (HCP). This plan protected 80% <strong>of</strong> remaining habitat on<br />

San Bruno Mountain, provided funds for the management and<br />

restoration <strong>of</strong> this habitat, and allowed for the development <strong>of</strong><br />

the remaining land. In 1982, the U.S. Congress institutionalised<br />

habitat conservation planning with amendments to the ESA,<br />

pointing to the Mission Blue conservation program as the<br />

model for this new process. Several dozen HCPs (many <strong>of</strong> them<br />

controversial) have been initiated in the ensuing decade.<br />

In a second noteworthy case, a similar impasse between<br />

developers and environmentalists has focused on the Karner<br />

Blue in upstate New York. As mentioned earlier, this highly<br />

threatened subspecies, protected by the state but not the U.S.<br />

government, is restricted to fire-maintained gaps in early<br />

successional pitch-pine and scrub-oak barrens. The rather<br />

sedentary blue exists as a limited suite <strong>of</strong> metapopulations,<br />

consisting <strong>of</strong> collections <strong>of</strong> local populations that are dependent<br />

on a shifting mosaic <strong>of</strong> suitable habitat.<br />

There has been an extensive, broad-based effort to conserve<br />

the remaining Karner Blue population (see review by Zaremba<br />

1991). Primarily through the joint efforts <strong>of</strong> The Nature<br />

Conservancy and the New York State Department <strong>of</strong><br />

Environmental <strong>Conservation</strong>, approximately 800ha <strong>of</strong> Karner<br />

Blue habitat have been protected as part <strong>of</strong> the Albany Pine<br />

Bush Preserve. In addition, the New York State Legislature<br />

established the Albany Bush Commission and charged the<br />

group with managing the remaining habitat. This and other<br />

ongoing programmes have involved prescribed burning,<br />

hostplant propagation, creation <strong>of</strong> effective dispersal corridors,<br />

and development <strong>of</strong> land-use practices to promote butterfly<br />

dispersal in areas adjacent to the preserve. Using the Karner<br />

Blue, Givnish et al. (1988) provided a model study for application<br />

<strong>of</strong> population viability analysis to conservation planning.<br />

Virtually identical in structure to an independently generated<br />

analysis <strong>of</strong> the well-studied nymphalid Euphydryas editha<br />

(Boisduval) (Murphy et al. 1990), this study broke from the<br />

traditional treatments <strong>of</strong> genetic threats and demographic<br />

stochasticity, and instead targeted environmental perturbations<br />

and metapopulation dynamics in an integrated scheme <strong>of</strong> reserve<br />

design and management.<br />

<strong>Conservation</strong> planning for lycaenids has focused primarily<br />

on habitat management. All <strong>of</strong> the listed taxa, and many <strong>of</strong> the<br />

candidates, survive as remnant populations in small 'garrison'<br />

reserves embedded within largely urban areas with long histories<br />

42<br />

<strong>of</strong> human settlement (Murphy 1987b). In such circumstances,<br />

opportunities to expand current distributions are few and<br />

conservation is less a matter <strong>of</strong> reserve design than reserve<br />

management. Heroic management efforts have brought Lange's<br />

Metalmark (Apodemia mormo langei Comstock) back from the<br />

brink <strong>of</strong> extinction. In 1976, the subspecies was listed as<br />

endangered by the U.S. government, as its remaining habitat<br />

was being threatened by sandmining and industrial development<br />

(see Opler 1991). The subspecies is restricted to the riparian<br />

sand dunes along the Sacramento and San Joaquin Rivers in<br />

central California. To prevent a further decline in numbers, in<br />

1980, the U.S. Fish and Wildlife Service acquired all <strong>of</strong> the<br />

lycaenid's remaining habitat. While seriously degraded by<br />

sandmining and invasive plants, this 24ha region was<br />

incorporated into the San Francisco National Wildlife Refuge,<br />

and there have been considerable efforts to restore the habitat.<br />

After a number <strong>of</strong> unsuccessful attempts to increase butterfly<br />

numbers, managers found that disking portions <strong>of</strong> the habitat<br />

resulted in dramatic growth <strong>of</strong> the larval and adult foodplant<br />

(Eriogonum nudutri). Since this action, the metalmark's<br />

abundance is estimated to have more than tripled, from fewer<br />

than 200 individuals in 1986 to more than 650 in 1989.<br />

Conclusions<br />

Extreme environmental events (such as drought, deluge,<br />

wildfire) can lead to dramatic fluctuations in the size <strong>of</strong> local<br />

butterfly populations, and in some well-documented cases, this<br />

has resulted in their extinction (e.g. Ehrlich et al. 1980; Murphy<br />

and Weiss 1988). For example, Ehrlich et al. (1972) reported<br />

that an early summer snowstorm caused the extinction <strong>of</strong> at<br />

least one subalpine population <strong>of</strong> Glaucopsyche lygdamus<br />

when it destroyed the entire standing crop <strong>of</strong> its larval food<br />

plant. Biotic factors, including the impact <strong>of</strong> natural enemies<br />

and ant associates, can also lead to significant variation in the<br />

size <strong>of</strong> lycaenid populations. However, as stressed throughout<br />

this volume, a long history <strong>of</strong> human-induced habitat alteration<br />

and destruction is responsible for the vast majority <strong>of</strong> declines<br />

and extinctions <strong>of</strong> lycaenids worldwide.<br />

This has certainly been the case for North American<br />

lycaenids, as already discussed for a number <strong>of</strong> taxa in this<br />

chapter. The El Segundo Blue (Euphilotes battoides allyni<br />

(Shields)) is yet another example. This small lycaenid is restricted<br />

to the El Segundo sand dunes along the coast <strong>of</strong> southern<br />

California. While most <strong>of</strong> its habitat, perhaps more than 10,000ha<br />

in extent, has been destroyed by housing and commercial<br />

developments, small portions <strong>of</strong> the sand-dune ecosystem (which<br />

support its preferred larval hostplant Eriogonum parvifolium)<br />

still remain near the Los Angeles International Airport and a<br />

Standard Oil refinery (Arnold 1983). While noteworthy efforts<br />

have been made by Standard Oil to establish populations <strong>of</strong> the<br />

El Segundo Blue on their lha property, comparatively little has<br />

been done by the City <strong>of</strong> Los Angeles on the inhabitable<br />

sections <strong>of</strong> their 80ha airport site. Since the mid 1970s, there


have been a number <strong>of</strong> unsuccessful attempts to develop much<br />

<strong>of</strong> the remaining site, several with revenue-generating activities<br />

linked to habitat management plans.<br />

Pressure to develop and disturb lycaenid habitat in North<br />

America is likely to intensify in coming decades, and more<br />

listings <strong>of</strong> endangered and threatened species should be expected.<br />

Only an increase in the already considerable efforts to conserve<br />

lycaenids (and insects in general) will suffice to keep more <strong>of</strong><br />

them from suffering the fate <strong>of</strong> the recently extirpated Palo<br />

Verdes Blue in southern California (see Arnold 1987).<br />

Establishment <strong>of</strong> the Xerces Society in 1971 has been an<br />

important step, as it has greatly increased the attention paid to<br />

insect conservation. The ESA <strong>of</strong> 1973 and the use <strong>of</strong> HCPs have<br />

also been instrumental in facilitating the protection <strong>of</strong> threatened<br />

and endangered lycaenids on both public and private lands.<br />

Unfortunately, the U.S. government has added only six insects<br />

to its list since 1981 (Opler 1991) – despite the fact that habitat<br />

degradation and the list <strong>of</strong> candidates have increased<br />

considerably during the intervening years. This reluctance to<br />

list species must be reversed, so that the power <strong>of</strong> the ESA and<br />

HCPs can have their designed effect (Murphy 1991).<br />

Although the ESA and HCPs are powerful instruments <strong>of</strong><br />

conservation planning and management, they are nevertheless<br />

'stop-gap' measures designed to take effect after taxa are in<br />

trouble. As attention continues to be focused on these essential<br />

management efforts, considerable effort must also be directed<br />

towards more long-term objectives associated with dramatically<br />

reducing the environmental degradation that is leading to the<br />

endangerment and extinction <strong>of</strong> additional taxa. As outlined by<br />

Ehrlich and Ehrlich (1990), Gore (1991) and others, these longterm<br />

objectives include stabilisation and then reduction <strong>of</strong> the<br />

growth <strong>of</strong> human populations, rapid development and<br />

deployment <strong>of</strong> environmentally appropriate technologies,<br />

comprehensive changes in the system <strong>of</strong> economic accounting<br />

so as to accurately reflect the effects <strong>of</strong> our actions on the<br />

environment, and the development <strong>of</strong> a detailed scheme for<br />

environmental education and research.<br />

References<br />

ARNOLD, R.A. 1983. Ecological studies <strong>of</strong> six endangered butterflies<br />

(Lepidoptera, <strong>Lycaenidae</strong>): island biogeography, patch dynamics, and the<br />

design <strong>of</strong> habitat preserves. Univ. Calif. Publns Entomol. 99: 1–161.<br />

ARNOLD, R.A.I 987. Decline <strong>of</strong> the endangered Palos Verdes blue butterfly<br />

in California. Biol. Conserv. 40: 203–217.<br />

ATSATT, P.R. 1981. Lycaenid butterflies and ants: selection for enemy-free<br />

space. Amer. Nat. 118: 638–654.<br />

BROWN, F.M. 1957. Colorado butterflies. Denver Museum <strong>of</strong> Natural<br />

History, Denver.<br />

BROWN, J.H. and KODRIC-BROWN, A. 1977. Turnover rates in insular<br />

biogeography: effect <strong>of</strong> immigration on extinction. Ecology 58: 445–449.<br />

COTTRELL, C.B. 1984. Aphytophagy in butterflies; its relationship to<br />

myrmecophily. Zool. J. Linn. Soc. 79: 1–57.<br />

DOWNEY, J.C. 1962. Myrmecophily in Plebejus(Icaricia) icarioides (Lepid.:<br />

<strong>Lycaenidae</strong>). Ent. News 73: 57–66.<br />

EHRLICH, P.R. 1965. The population biology <strong>of</strong> the butterfly, Euphydryas<br />

editha. II. The structure <strong>of</strong> the Jasper Ridge colony. Evolution 19:327–336.<br />

43<br />

EHRLICH, P.R., BREEDLOVE, D.E., BRUSSARD, P.E. and SHARP, M.A.<br />

1972. Weather and the 'regulation' <strong>of</strong> subalpine populations. Ecology 53:<br />

243–247.<br />

EHRLICH, P.R. and EHRLICH, A.H. 1990. The population explosion. Simon<br />

and Schuster, New York.<br />

EHRLICH, P.R., MURPHY, D.D., SINGER, M.C., SHERWOOD, C.B.,<br />

WHITE, R.R. and BROWN, I.L. 1980. Extinction, reduction, stability and<br />

increase: the response <strong>of</strong> Checkerspot butterfly (Euphydryas) populations<br />

to the California drought. Oecologia 46: 101–105.<br />

EMMEL, T.C. and EMMEL, J.F. 1973. The <strong>Butterflies</strong> <strong>of</strong> Southern California.<br />

Natural History Museum <strong>of</strong> Los Angeles County, Los Angeles.<br />

FEDERAL REGISTER 1991a. Endangered and Threatened Wildlife and<br />

Plants. U.S. Department <strong>of</strong> the Interior, Fish and Wildlife Service (15 July<br />

1991).<br />

FEDERAL REGISTER 1991b. Endangered and threatened wildlife and<br />

plants; animal candidate review for listing as endangered or threatened<br />

species, proposed rule. U.S. Department <strong>of</strong> the Interior, Fish and Wildlife<br />

Service (21 November 1991).<br />

GIVNISH, T., MENGES, E and SCHWEITZER, D. 1988. Minimum area<br />

requirements for long-term conservation <strong>of</strong> the Albany Pine Bush and<br />

Karner blue butterfly (Volume IV, Appendix T). Report to the City <strong>of</strong><br />

Albany, Malcolm Pirnie, Inc.<br />

GORE, A. 1991. Earth in the Balance: Ecology and the Human Spirit.<br />

Houghton Mifflin, New York.<br />

HARRIS, L. 1972. <strong>Butterflies</strong> <strong>of</strong> Georgia. University <strong>of</strong> Oklahoma Press,<br />

Norman.<br />

MURPHY, D.D. 1987a. A report <strong>of</strong> the California butterflies listed as<br />

candidates for endangered status by the United States Fish and Wildlife<br />

Service. Report C-1755 to the California Department <strong>of</strong> Fish and Game.<br />

MURPHY, D.D. 1987b. Challenges to biological diversity in urban areas. In:<br />

Wilson, E.O., (Ed.) Biodiversity. National Academy <strong>of</strong> Sciences Press,<br />

Washington, D.C., pp. 71–76.<br />

MURPHY, D.D. 1991. Invertebrate conservation. In: Kohm, K.A. (Ed.)<br />

Balancing on the Brink <strong>of</strong> Extinction: the Endangered Species Act and<br />

Lessons for the Future. Island Press, Washington, D.C., pp. 181–198.<br />

MURPHY, D.D., FREAS, K.E. and WEISS, S.B. 1990. An environmentmetapopulation<br />

approach to population viability analysis for a threatened<br />

invertebrate. Conserv. Biol. 4: 41–51.<br />

MURPHY, D.D., MENNINGER, M.S., EHRLICH, P.R. and WILCOX, B.A.<br />

1986. Local population dynamics <strong>of</strong> adult butterflies and the conservation<br />

status <strong>of</strong> two closely related species. Biol. Conserv. 37: 201–223.<br />

MURPHY, D.D. and WEISS, S.B. 1988. Ecological studies and conservation<br />

<strong>of</strong> the Bay Checkerspot butterfly, Euphydryas editha bayensis. Biol.<br />

Conserv. 46: 183–200.<br />

NEW.T.R. 1991. Butterfly <strong>Conservation</strong>. Oxford University Press, Melbourne.<br />

OPLER, P.A. 1991. North American problems and perspectives in insect<br />

conservation. In: Collins, N.M. and Thomas, J. A. (Eds). The <strong>Conservation</strong><br />

<strong>of</strong> Insects and their Habitats. Academic Press, New York, pp. 9–32.<br />

ORSAK, L.J. 1978. The <strong>Butterflies</strong> <strong>of</strong> Orange County, California. University<br />

<strong>of</strong> California Press, Irvine.<br />

PIERCE, N.E. 1987. The evolution and biogeography <strong>of</strong> associations between<br />

lycaenid butterflies and ants. In: Harvey, P.H. and Partridge, L. (Eds).<br />

Oxford Surveys in Evolutionary <strong>Biology</strong>, Volume 4. Oxford University<br />

Press, Oxford, pp. 89–116.<br />

PIERCE, N.E. and EASTEAL, S. 1986. The selective advantage <strong>of</strong> attendant<br />

ants from the larvae <strong>of</strong> a lycaenid butterfly, Glaucopsyche lygdamus. J.<br />

Anim. Ecol. 55: 451–462.<br />

PIERCE, N.E. and MEAD, P.S. 1981. Parasitoids as selective agents in the<br />

symbiosis between lycaenid butterfly larvae and ants. Science 211:<br />

1185–1187.<br />

PRATT, G.F. and BALLMER, G.R. 1986. The phenetics and comparative<br />

biology <strong>of</strong> Euphilotes enoptes from the San Bernardino Mountains.J. Res.<br />

Lepid. 25: 121–135.<br />

RAUSHER, M.D. 1982. Population differentiation in Euphydryas editha<br />

butterflies: larval adaptations to different host plants. Evolution 36:<br />

581–590.<br />

REID, T.S. and MURPHY, D.D. 1986. The endangered mission blue butterfly,


Plebejus icarioides missionensis. In: Wilcox, B.A., Brussard, P.F. and<br />

Marcot, B.G. (Eds). The Management <strong>of</strong> Viable Populations: Theory,<br />

Application, and Case Studies. Center for <strong>Conservation</strong> <strong>Biology</strong>, Stanford,<br />

California, pp. 147–168.<br />

SCOTT, J.A. 1986. The <strong>Butterflies</strong> <strong>of</strong> North America. Stanford University<br />

Press, Stanford, California.<br />

SINGER, M.C. 1971. Evolution <strong>of</strong> food-plant preferences in the butterfly<br />

Euphydryas editha. Evolution 35: 383–389.<br />

SINGER, M.C. 1983. Determinants <strong>of</strong> multiple host use by a phytophagous<br />

insect population. Evolution 37: 389–403.<br />

SINGER, M.C, NG, D. and THOMAS, CD. 1988. Heritability <strong>of</strong> oviposition<br />

preference and its relationship to <strong>of</strong>fspring performance within a single<br />

insect population. Evolution 42: 977–985.<br />

THOMAS, J.A. 1980. Why did the Large Blue become extinct in Britain?<br />

Oryx 15: 243–247.<br />

TILDEN, J.W. 1965. <strong>Butterflies</strong> <strong>of</strong> the San Francisco Bay region. University<br />

<strong>of</strong> California Press, Berkeley, California.<br />

VRIJENHOEK, R.C 1985. Animal population genetics and disturbance: the<br />

44<br />

effects <strong>of</strong> local extinctions and recolonisation on heterozygosity and<br />

fitness. In: Pickett, S.T.A. and White, P.S. (Eds). The Ecology <strong>of</strong> Natural<br />

Disturbance and Patch Dynamics. Academic Press, New York, pp. 265–285.<br />

WEBSTER, R.P. and NIELSON, M.C. 1984. Myrmecophily in the Edward's<br />

hairstreak butterfly Satyrium edwardsii (<strong>Lycaenidae</strong>). J. Lepid. Soc. 38:<br />

124–133.<br />

WEISS, S.B. and MURPHY, D.D. 1990. Thermal microenvironments and the<br />

restoration <strong>of</strong> rare butterfly habitat. In: Berger, J. (Ed). Environmental<br />

Restoration: Sciences and Strategies for Restoring the Earth. Island Press,<br />

Washington, D.C., pp. 55–60.<br />

WHITE, R.R. and SINGER, M.C. 1974. Geographical distribution <strong>of</strong> hostplant<br />

choice in Euphydryas editha (Nymphalidae). J. Lepid. Soc. 28: 103–107.<br />

ZAREMBA, R.E. 1991. Management <strong>of</strong> Karner blue butterfly habitat in a<br />

suburban landscape. In: Decker, D.J., Krasny, M.E., G<strong>of</strong>f, G.R., Smith,<br />

C.R. and Gross, D.W. (Eds). Challenges in the <strong>Conservation</strong> <strong>of</strong> Biological<br />

Resources: a Practitioner's Guide. Westview Press, San Francisco, pp.<br />

289–302.


Neotropical <strong>Lycaenidae</strong>: an overview<br />

Keith S. BROWN, JR.<br />

Departamento de Zoologia, Instituto de Biologia, Universidade Estadual de Campinas, C.P. 6109 Campinas,<br />

Sao Paulo 13.081, Brazil<br />

Introduction<br />

The neotropical <strong>Lycaenidae</strong> are still only partly known and<br />

very little studied. They are here taken to include the Riodininae<br />

which are possibly closer to Nymphalidae than to other Lycaenids<br />

(Robbins 1988).The total <strong>of</strong> approximately 2300 species (Tables<br />

1–3; Robbins 1982,1992; Harvey 1987; Callaghan and Lamas<br />

in press) includes as many as 400 still to be described, and<br />

probably an equal number <strong>of</strong> well-known names which will be<br />

synonymised or joined with others as subspecies. With the<br />

exceptions <strong>of</strong> a single copper and about 60 blues, the species are<br />

divided almost equally between hairstreaks (all in the single<br />

tribe Eumaeini <strong>of</strong> the subfamily Theclinae) and metalmarks<br />

(Riodininae, with four neotropical tribes containing 1, 1, 137<br />

and about 1100 named species; the last tribe is separable into at<br />

least eight subtribes (Harvey 1987)).<br />

The neotropical fauna includes some <strong>of</strong> the most exquisite<br />

colours and bizarre patterns known in butterflies (Figure 1). It<br />

is perhaps fortunate that they attract little attention from collectors<br />

and dealers although as a result <strong>of</strong> this the body <strong>of</strong> biological<br />

and distributional information available is far less than for the<br />

swallowtails (Papilionidae, see Collins and Morris 1985) or<br />

some groups <strong>of</strong> Nymphalidae. A reasonable cross-section <strong>of</strong><br />

phenotypes can be seen in colour in Hewitson's original<br />

descriptions and illustrations (1852–1872, 1863–1878), in<br />

Barcant's book on Trinidad butterflies (1970, Plates 5, 9, 11,<br />

12, 27, 28 plus 22 and 23 in black and white), in Lewis's<br />

'<strong>Butterflies</strong> <strong>of</strong> the World' (1973), and in Seitz'<br />

'Macrolepidoptera <strong>of</strong> the World' (1916–1920, Plates 121–159,<br />

110A, 113B, 193) but with many names outdated. Only minimal<br />

information can be unearthed in most general butterfly books<br />

(Smart, 1975, illustrates only 103 species) or local lists for the<br />

neotropics (de la Maza, 1988, shows only 118 species). The<br />

endemic Chilean lycaenid fauna is ignored in most publications,<br />

and many genera <strong>of</strong> South American hairstreaks have no name<br />

as yet.<br />

Less than 20% <strong>of</strong> the neotropical Theclinae (Robbins 1993)<br />

and 10% <strong>of</strong> the Riodininae (Harvey 1987) have been subject to<br />

any biological study (e.g. population censuses, juvenile biology<br />

and host plants, myrmecophily, behaviour, voltinism). While a<br />

45<br />

few species feed on plants <strong>of</strong> economic importance, especially<br />

Orchidaceae, Bromeliaceae, Leguminosae, Sapotaceae,<br />

Solanaceae, Myrtaceae, Anacardiaceae, Rubiaceae and<br />

Compositae, even these are rarely reared by entomologists.<br />

Highly diversified faunas in the high mountains (Theclinae)<br />

and lowland Amazonia (Riodininae) are so poorly sampled that<br />

even fundamental questions – on species relationships, generic<br />

assignments, distributions, resource partitioning, optional or<br />

obligatory relationships with ants, association <strong>of</strong> dimorphic<br />

females with their males, flight habits, seasonal variation in<br />

pattern and abundance, migration – remain to be answered.<br />

In such a frame, the picture is ill-defined, begging for much<br />

new work by biologists who are not averse to meeting the small,<br />

the little-known, the variable and complex; all too few have<br />

accepted this challenge. Although it may be valid that 'A true<br />

connoisseur <strong>of</strong> neotropical Lepidoptera can always be<br />

distinguished by his love for the Lycaenids' (Brown 1973), this<br />

is still an affair destined for frustration.<br />

Thus, the following attempts at generalisation and<br />

particularisation are very fragile, begging for more field work,<br />

laboratory study and experimentation. Very preliminary answers<br />

can be attempted for the following questions:<br />

(1) Are neotropical <strong>Lycaenidae</strong> 'typical' members <strong>of</strong> the<br />

family, or do they show their own biological styles and<br />

syndromes?<br />

(2) Do neotropical <strong>Lycaenidae</strong> show clear patterns <strong>of</strong><br />

distribution, variation, behaviour, community structure, and<br />

ecological interaction?<br />

(3) Are neotropical <strong>Lycaenidae</strong> useful as indicators <strong>of</strong> other<br />

animal and plant species, historical and ecological factors,<br />

system characteristics including degree <strong>of</strong> disturbance, and<br />

general community structure and function?<br />

(4) Is it possible to identify threatened species or groups? Do<br />

they co-occur with endangered communities <strong>of</strong> other animals<br />

and plants? Can they be saved?<br />

Most <strong>of</strong> the answers must be sought by patient and diligent<br />

field work in the neotropics. In the past few years, a few<br />

scientists have come to terms with the systematic tangles <strong>of</strong> the<br />

group and have begun to do careful studies <strong>of</strong> the biology <strong>of</strong><br />

some species. Hopefully, their spectacular and fascinating<br />

results (see De Vries 1990, 1991a) will attract more workers to


the family, to learn more about these small but disproportionately<br />

impressive, beautiful and varied insects (Figure 1).<br />

Systematics and ecology<br />

A summary <strong>of</strong> the recognized divisions <strong>of</strong> neotropical<br />

<strong>Lycaenidae</strong> down to the generic level (partial for the Theclinae)<br />

is presented in Tables 1–3, along with distributional, biological<br />

and bibliographic data. A number <strong>of</strong> outstanding and salient<br />

46<br />

facts characterise the neotropical <strong>Lycaenidae</strong>:<br />

• the small number <strong>of</strong> blues (Polyommatinae) in the region,<br />

with very widespread ocurrence <strong>of</strong> only three common<br />

species;<br />

• the predominance <strong>of</strong> forest groups, with very few species<br />

dependent on non-forest, open or successional habitats as is<br />

more common in the northern hemisphere;<br />

• the relatively low proportion <strong>of</strong> myrmecophilous species,<br />

and almost complete absence <strong>of</strong> lichen or fungus-eating<br />

species (although Calycopis larvae may <strong>of</strong>ten be detritivorous<br />

(Johnson 1985; Robbins 1992) and Sarota larvae feed on<br />

Figure 1. A pot-pourri <strong>of</strong> lycaenid diversity in the neotropics, all from<br />

Japi, São Paulo: (a) Theclinae.<br />

Key to names:<br />

1 'Thecla' phydela; 2 Atlides polybe; 3 Evenus regalis; 4 Arcas ducalis;<br />

5 Panthiadesphaleros; 6 Erora campa; 7 Chalybs Hassan; 8 Chalybs chloris;<br />

9 Arawacus tarania; 10 Cyanophrys acaste; 11 Cyanophrys bertha;<br />

12 Cyanophrys remits; 13 Brangas ca. didymon; 14 Chlorostrymon simaethis;<br />

15 Erora ca. opisena; 16 Ocaria cinerea; 17 'Thecla' deniva; 18 Ipidecla<br />

schausi; 19 Rekoa melon; 20 Ministrymon No.l.; 21 Rekoa meton Ventral;<br />

22 Brangas silumena; 23 Ocaria thales; 24 Magnastigma hirsuta;<br />

25 Chlorostrymon telea; 26 Parrhasius orgia; 27 'Thecla' elika; 28<br />

Ministrymon No. 2.; 29 Thereus cithonius; 30 Parrhasius selika. Ventral side<br />

shown, except 19, 25, 26, 27, 28 dorsal.


epiphylls – liverworts and blue-green algae (De Vries<br />

1988a)), and with only one case <strong>of</strong> carnivorous larvae<br />

known to date;<br />

• the rarified distribution <strong>of</strong> most species – widespread but<br />

very sporadic;<br />

• the small number <strong>of</strong> serious pest species considering the<br />

wide range <strong>of</strong> host plants;<br />

• the presence <strong>of</strong> two monotypic tribes (Stygini and<br />

Corrachiini) in the Riodininae;<br />

• the large number <strong>of</strong> monotypic genera and the existence <strong>of</strong><br />

a few gargantuan genera (Strymon, Calycopis, Euselasia,<br />

47<br />

Mesosemia) (both perhaps due to insufficient study and the<br />

complexity <strong>of</strong> the family).<br />

Some <strong>of</strong> these patterns may be altered when the groups are<br />

better known, but at the moment the 'flavour' is strongly that <strong>of</strong><br />

typical forest butterflies. This may be due to the predominance<br />

<strong>of</strong> forest biomes in the neotropics, but even open-vegetation<br />

genera (Rekoa, Strymon, Electrostrymon, Chlorostrymon,<br />

Calephelis, Ematurgina, Audre, Apodemia, Lemonias, Aricoris)<br />

show biological syndromes much like those <strong>of</strong> their forestinhabiting<br />

relatives and in contrast with other <strong>Lycaenidae</strong><br />

which characterise successional habitats.<br />

Figure 1. A pot-pourri <strong>of</strong> lycaenid diversity in the neotropics, all from<br />

Japi, São<br />

Paulo: (b) Riodininae and Polyommatinae.<br />

Key to names:<br />

1 Lemonias glaphyra; 2 Calydna sp. n. nr. hemis; 3 Xenandra heliodes;<br />

4 Symmachia arion Female; 5 'Mesosemia' acuta; 6 Mesene pyrippe;<br />

7 Anteros lectabilis; 8 Napaea phryxe; 9 Emesis fastidiosa Male Ventral;<br />

10 Chorinea licursis; 11 Emesis fastiosa Female Ventral; 12 Barbicornis<br />

basilis; 13 Notheme erota; 14 Charts cadytis; 15 Lasaia agesilas; 16 Caria<br />

plutargus; 17 Synargis brennus; 18 Calephelis brasiliensis; 19 Pterographium<br />

sagaris satnius; 20 'Everes' cogina Female; 21 Zizula cyna tultiola;<br />

22 Baeotus johannae; 23 Adelolypa bolena; 24 Emesis fatimella; 25 Parcella<br />

amarynthina Female; 26 Theope thestias ca. discus; 27 Leucochimona matalha;<br />

28 Panara soana trabalis; 29 Lemonias zygia epona; 30 Eurybia pergaea var.;<br />

31 Riodina lycisca; 32 Mesosemia odice Female. Dorsal aspect shown, except<br />

1, 2, 7, 8, 9, 11, 14, 18, 24, 28, 32 ventral.


Table 1. Synopsis <strong>of</strong> neotropical Riodininae a ; systematics and biology b .<br />

Taxonomic groups:<br />

SUBFAMILY (TRIBE),<br />

Sublribe or Group,<br />

"Genus;<br />

RIODININAE g (STYGINI)<br />

S,yx<br />

(CORRACHIINI)<br />

Corrachia<br />

(EUSALASHNI)<br />

* † Euselasia<br />

Hades<br />

Methone<br />

(RIODININI)<br />

Mesosemiiti h<br />

Pcrophlhalma<br />

Mesophthalma<br />

*Leucochimona<br />

*Semomesia<br />

* † Mesosemia<br />

* † Eunogyra<br />

*Eurybia<br />

†<br />

Aleesa<br />

Mimocaslnia<br />

* Teralophthalma<br />

*Ithomiola<br />

Vollinia<br />

Hyphilaria<br />

Hermathena<br />

Cremna<br />

* † Napaea<br />

Eucorna<br />

Riodiniti i,p<br />

*Lyropleryx<br />

Necyria<br />

Cyrenia<br />

*Ancyluris<br />

Nirodia j<br />

Rhetus<br />

Chorinea<br />

'Nahida<br />

*lthomeis<br />

*Panara<br />

Isapis<br />

*Brachyglenis<br />

*Themone<br />

Plate<br />

numbers<br />

(Lewis<br />

1973)<br />

78:33<br />

71:42<br />

73:4–32<br />

74:1–2<br />

76:3<br />

77:22<br />

75:16<br />

74:24<br />

78:19<br />

72:4–5<br />

75:14–37<br />

72:29<br />

72:30–4<br />

73:1–3<br />

70:1<br />

76:4<br />

79:7–8<br />

74:12–3<br />

79:31<br />

74:5–7<br />

74:9<br />

71:43<br />

76:7–8<br />

:10–12<br />

–<br />

75:1–2<br />

76:5<br />

:13–4<br />

72:1<br />

70:4–12<br />

–<br />

78:8,12<br />

71:37<br />

76:6<br />

74:10–1<br />

77:18<br />

75:3<br />

71:2<br />

79:10<br />

Approx.<br />

no. <strong>of</strong><br />

spp.<br />

1<br />

1<br />

134<br />

2<br />

1<br />

1<br />

1<br />

9<br />

8<br />

120<br />

1<br />

20<br />

6<br />

1<br />

6<br />

3<br />

2<br />

6<br />

2<br />

5<br />

13<br />

1<br />

4<br />

6<br />

1<br />

21<br />

1<br />

3<br />

7<br />

4<br />

9<br />

6<br />

1<br />

5<br />

4<br />

Typical or<br />

well-known<br />

species<br />

infernalis<br />

leucoplaga<br />

mys, geon<br />

noctula<br />

Cecilia<br />

tullius<br />

idotea<br />

mathata<br />

capanea<br />

cippus,<br />

telegone<br />

satyrus<br />

nicaea<br />

prema, amesis<br />

rothschildi<br />

phelina<br />

floralis<br />

radiata<br />

nicia<br />

candidata<br />

actoris, thasus<br />

eucharila<br />

sanarila<br />

apollonia<br />

bellona<br />

martia<br />

aulestes<br />

belphegor<br />

periander<br />

octauius<br />

coenoides<br />

astraea<br />

episatnius<br />

agyrtus<br />

esthema<br />

pais, poecila<br />

Distribution<br />

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

nouns c<br />

Peru And<br />

CR<br />

NT, sp.AM<br />

TRAn<br />

AM–CR<br />

NT<br />

AM<br />

NT<br />

AM–BA<br />

NT<br />

AM-BA<br />

NT<br />

SAm<br />

AM<br />

AM–And<br />

AM<br />

TRAn<br />

SAm<br />

NT<br />

NT<br />

NT<br />

BR–SM<br />

NT<br />

And–CR<br />

AM<br />

NT<br />

BR–MG<br />

NT<br />

NT<br />

EcAnd<br />

AM–CR<br />

AM All<br />

NT<br />

NT<br />

AM<br />

48<br />

Mim d<br />

?<br />

– ?<br />

+–<br />

+<br />

++<br />

–<br />

–<br />

+<br />

–<br />

_<br />

–<br />

_<br />

–<br />

–<br />

+–?<br />

++<br />

–<br />

+–<br />

+?<br />

–<br />

_<br />

–<br />

+<br />

+?<br />

–<br />

+–<br />

–<br />

–<br />

+<br />

++<br />

++<br />

+–<br />

–<br />

++<br />

++<br />

Myr e<br />

–<br />

–<br />

–<br />

–<br />

–<br />

–<br />

–<br />

–<br />

_<br />

+–?<br />

+<br />

+<br />

+<br />

–<br />

?<br />

–<br />

–<br />

_<br />

?<br />

–?<br />

–?<br />

–?<br />

–<br />

–?<br />

–<br />

_<br />

–?<br />

–?<br />

–?<br />

–?<br />

–?<br />

–?<br />

Time<br />

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

activity<br />

PM<br />

?<br />

AM.PM<br />

AM<br />

AM<br />

PM<br />

MD<br />

PM<br />

PM<br />

AM.PM<br />

AM<br />

LPM<br />

MD<br />

7<br />

PM<br />

MD<br />

LPM<br />

AM<br />

PM<br />

PM<br />

PM<br />

PM<br />

MD<br />

MD<br />

MD<br />

AM<br />

MD<br />

AM.MD<br />

MD<br />

?<br />

PM<br />

PM<br />

AM<br />

AM<br />

PM<br />

Habitat<br />

(usual) c<br />

CloudF<br />

CloudF<br />

HumidF<br />

RainF<br />

RainF<br />

HumidF<br />

HumidF<br />

HumidF<br />

RainF<br />

HumidF<br />

HumidF<br />

RiverF<br />

RainF<br />

RainF<br />

CloudF<br />

HumidF<br />

CloudF<br />

HumidF<br />

OpenF<br />

RainF<br />

HumidF<br />

CloudF<br />

RainF.Cd<br />

CloudF<br />

HumidF<br />

HumidF<br />

CmpRup<br />

HumidF<br />

HumidF<br />

CloudF<br />

RainF<br />

RainF<br />

RainF<br />

HumidF<br />

RainF<br />

Abn f<br />

1<br />

1<br />

2–5<br />

3<br />

2<br />

3<br />

2<br />

4<br />

3<br />

2–5<br />

3<br />

4<br />

2<br />

1<br />

3<br />

3<br />

2<br />

3<br />

2<br />

3–4<br />

3<br />

1<br />

2<br />

4<br />

2<br />

3<br />

2<br />

4<br />

3<br />

2<br />

3<br />

3<br />

3<br />

2<br />

2<br />

Larval<br />

host<br />

plants')<br />

Unknown<br />

Unknown<br />

Myrt, Clusi.<br />

Anacardi.<br />

Unknown<br />

Rubi.<br />

?<br />

Rubi.<br />

?<br />

Rubi.<br />

Marant.<br />

Solan.<br />

?<br />

?<br />

?<br />

?<br />

Orchid.<br />

?<br />

?<br />

Orchid., Bromeli.<br />

?<br />

?<br />

?<br />

?<br />

Melastomat.<br />

?<br />

?<br />

Flacourti.,<br />

Celastr.<br />

?<br />

?<br />

?<br />

?<br />

?<br />

?<br />

Bibliography<br />

1,2,3,23<br />

1<br />

23<br />

23<br />

23<br />

4<br />

23<br />

23<br />

23<br />

23


Table 1 (cont). Synopsis <strong>of</strong> neotropical Riodininae a ; systematics and biology b .<br />

Taxonomic groups:<br />

SUBFAMILY (TRIBE),<br />

Subtribe or Group,<br />

* † Genus;<br />

Notheme<br />

Monethe<br />

Paraphthonia<br />

†<br />

Colaciticus<br />

Meiacharis<br />

†<br />

Cariomothus<br />

* † Lepricornis<br />

Pheles<br />

Barbicornis<br />

Syrmatia<br />

*Chamaelimnas<br />

Carlea<br />

Crocozona<br />

† *Baeotis<br />

Caria<br />

"Chalodeta<br />

Parcel/a<br />

† *Charts<br />

*Calephelis<br />

Amarynthis<br />

Amphiselenis<br />

* † Lasuia<br />

†<br />

Exoplisia<br />

Riodina<br />

* † Melanis<br />

:29–32<br />

* † Siseme<br />

†<br />

Comphotis<br />

Symmachiiti<br />

Luciliella<br />

*Mesene<br />

:10–4<br />

†<br />

Mesenopsis<br />

* † Xenandra<br />

79:35,38<br />

†<br />

Xynias<br />

* † Esthemopsis<br />

Chimaslrum<br />

† *Symmachia<br />

†<br />

*Pterographium<br />

†<br />

*Phaenochitonia<br />

Plate<br />

numbers<br />

(Lewis<br />

1973)<br />

76:19<br />

76:9<br />

–<br />

71:39<br />

76:1–2<br />

71:19–20<br />

74:23<br />

77:29<br />

70:38<br />

71:1<br />

79:9<br />

71:25–8<br />

71:21<br />

71:44<br />

70:23<br />

:35–7<br />

71:15–8<br />

71:22–3<br />

:34<br />

77:20<br />

72:23<br />

71:29–31<br />

:33<br />

(20:45–6)<br />

70:2<br />

70:3<br />

74:14–6<br />

76:15–7<br />

78:13–5<br />

74:26–7<br />

78:24–8<br />

71:41<br />

74:25<br />

75:5–6<br />

75:15<br />

74:28<br />

79:36<br />

72:25–8<br />

71:36<br />

78:10–1<br />

:22,31<br />

:34<br />

79:1–6<br />

78:21<br />

77:25–6<br />

Approx.<br />

no. <strong>of</strong><br />

spp.<br />

1<br />

3<br />

2<br />

2<br />

9<br />

4<br />

11<br />

1<br />

1 k<br />

4<br />

11<br />

1<br />

4<br />

14<br />

14<br />

9<br />

1<br />

15<br />

32<br />

1<br />

1<br />

11<br />

4<br />

3<br />

39<br />

10<br />

3<br />

3<br />

28<br />

3<br />

9<br />

4<br />

14<br />

1<br />

45<br />

9<br />

12<br />

Typical or<br />

well-known<br />

species<br />

erota<br />

alphonsus<br />

mitlone<br />

johnstoni<br />

ptolemaeus<br />

erythromelas<br />

atricolor<br />

heliconides<br />

basilis<br />

nyx, aethiops<br />

tircis, briola<br />

vilula<br />

caecius<br />

hisbon, zonata<br />

ino, trochilus<br />

jessa, theodora<br />

amarynthina<br />

auius, cleonus<br />

nilus<br />

meneria<br />

chama<br />

agesilas<br />

cadmeis<br />

lycisca<br />

xarife, pixe<br />

arisioteles<br />

irrorata<br />

camissa<br />

phareus<br />

bryaxis<br />

heliodes<br />

cynosema<br />

inaria<br />

argenteum<br />

probetor<br />

sagaris<br />

cingulus<br />

Distri- Mimd<br />

bution<br />

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

genus c<br />

NT –<br />

NT –<br />

Peru ?<br />

AM –<br />

NT –<br />

SAm –<br />

SAm +?<br />

SAm +<br />

All +–<br />

SAm +–<br />

NT +<br />

AM<br />

UpAM, –<br />

BR–SM<br />

++<br />

NT +<br />

NT –<br />

NT –<br />

NT –<br />

NT –<br />

NT(+NA) –<br />

AM –<br />

Venez –<br />

NT –<br />

SAm –<br />

SAm –<br />

NT +–<br />

And –<br />

AM –<br />

TRAn –<br />

NT –<br />

NT +<br />

SAm –<br />

AM ++<br />

NT ++<br />

TRAn +–?<br />

NT +–<br />

SAm –<br />

NT –<br />

49<br />

Myr e<br />

–?<br />

–<br />

–<br />

–<br />

–?<br />

–<br />

–?<br />

–<br />

–<br />

–<br />

–?<br />

–?<br />

–<br />

–<br />

–?<br />

_<br />

–<br />

–?<br />

–?<br />

–?<br />

–?<br />

–?<br />

–<br />

–?<br />

–?<br />

–?<br />

–<br />

–<br />

–?<br />

–<br />

–<br />

–<br />

–<br />

Time<br />

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

activity<br />

AM<br />

MD<br />

?<br />

PM<br />

AM,PM<br />

PM<br />

AM<br />

MD<br />

AM<br />

EAM<br />

AM<br />

PM<br />

AM<br />

AM<br />

MD<br />

MD<br />

AM<br />

MD<br />

MD<br />

MD<br />

MD<br />

MD<br />

MD<br />

MD<br />

MD<br />

MD<br />

PM<br />

PM<br />

PM<br />

PM<br />

MD<br />

PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

Habitat<br />

(usual) c<br />

HumidF<br />

HumidF<br />

?<br />

RainF<br />

HumidF<br />

HumidF<br />

HumidF<br />

HumidF<br />

HumidF<br />

RainF<br />

HumidF<br />

RainF<br />

HumidF<br />

F, Cd<br />

HumidF<br />

HumidF<br />

HumidF<br />

HumidF<br />

Cmp<br />

HumidF<br />

CloudF<br />

RiverF<br />

RainF<br />

RiverF<br />

HumidF<br />

RiverF<br />

RainF<br />

CloudF<br />

HumidF<br />

RainF<br />

RainF<br />

RainF<br />

RainF<br />

RainF<br />

RainF<br />

HumidF<br />

HumidF<br />

Abn f<br />

3<br />

3<br />

?<br />

1<br />

4<br />

2<br />

2<br />

2<br />

4<br />

4<br />

2<br />

4<br />

3<br />

2<br />

4<br />

3<br />

3<br />

5<br />

5<br />

4<br />

?<br />

4<br />

2<br />

4<br />

4<br />

4<br />

2<br />

2<br />

3<br />

1<br />

1<br />

1<br />

1<br />

2<br />

1–2<br />

3<br />

3<br />

Larval Bibliohost<br />

graphy<br />

plants q<br />

?<br />

?<br />

?<br />

?<br />

Flacourti., 5<br />

Loranth.<br />

?<br />

Combret.<br />

?<br />

Sapor., Ulm. 6<br />

Zinziber<br />

?<br />

?<br />

?<br />

?<br />

Ulm. 2,3,7<br />

Sterculi., Aster.<br />

?<br />

?<br />

Aster. 2,3,8<br />

?<br />

?<br />

?<br />

?<br />

? 23<br />

Legumin., Aster. 3<br />

?<br />

?<br />

?<br />

Sapind.<br />

?<br />

?<br />

?<br />

?<br />

?<br />

?<br />

Melastom. 5,23<br />

?


Table 1 (cont). Synopsis <strong>of</strong> neotropical Riodininae a ; systematics and biology b .<br />

Taxonomic groups:<br />

SUBFAMILY (TRIBE),<br />

Subtribe or Group,<br />

* † Genus;<br />

† *Stichelia<br />

Helicopiti 1<br />

†<br />

*Sarota<br />

†<br />

*Anteros<br />

Ourocnemis<br />

Helicopis m<br />

Emesiti<br />

* † Argyrogrammana<br />

Callistium<br />

† *Calydna<br />

* † Emesis<br />

Pixus<br />

† *Pachythone<br />

Pseudonymphidia<br />

Roeberella<br />

l.amphiotes<br />

Apodemia<br />

Zabuella<br />

Dinoplolis<br />

Echenais<br />

Imelda<br />

Astraeodes<br />

Dianesia<br />

Mycasior<br />

Petrocerus<br />

Lemoniiti n<br />

*L.emonias<br />

Thisbe<br />

Uraneis<br />

Catocyclolis<br />

* † Juditha<br />

* † Synargis<br />

Thyranota<br />

* † Ematurgina<br />

† *Audre<br />

†<br />

Aricoris<br />

Eiseleia<br />

Plate<br />

numbers<br />

(Lewis<br />

1973)<br />

77:24,28<br />

78:18<br />

71:32–3<br />

70:13–8<br />

77:14<br />

74:3–4<br />

70:21–2<br />

:24–5<br />

:27<br />

71:5<br />

71:7–14<br />

:38,40<br />

72:18–22<br />

:24<br />

–<br />

77:15–7<br />

–<br />

78:17<br />

–<br />

70:20?<br />

79:33<br />

72:2<br />

–<br />

74:8<br />

70:26<br />

70:19<br />

–<br />

74:19–20<br />

:22<br />

79:27–30<br />

79:32,37<br />

71:23<br />

74:17–8<br />

78:7<br />

72:17<br />

76:26–31<br />

77:1–4<br />

79:34<br />

72:14<br />

70:28–34<br />

77:13<br />

–<br />

Approx.<br />

no. <strong>of</strong><br />

spp.<br />

7<br />

11<br />

14<br />

2<br />

4<br />

18<br />

1<br />

25<br />

46<br />

1<br />

14<br />

1<br />

3<br />

1<br />

12<br />

1<br />

1<br />

1<br />

3<br />

1<br />

1<br />

3<br />

1<br />

10<br />

4<br />

3<br />

1<br />

6<br />

26<br />

1<br />

5<br />

22<br />

2<br />

1<br />

Typical or<br />

well-known<br />

species<br />

bocchoris<br />

gyas, chrysus<br />

formosus<br />

archytas<br />

cupido<br />

holosticta<br />

cleadas<br />

thersander<br />

cerea, lucinda<br />

corculum<br />

gigas<br />

clearista<br />

calvus<br />

velazquesi<br />

mormo<br />

tenella<br />

orphana<br />

thelephus<br />

mycea<br />

areuta<br />

carteri<br />

leucarpis<br />

catiena<br />

zygia<br />

irenea, molela<br />

hyalina<br />

aemulius<br />

azan, molpe<br />

tytia, abaris<br />

galena<br />

axenus<br />

epulus<br />

tutana<br />

terias<br />

Distribution<br />

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

genus c<br />

SAm<br />

NT<br />

NT<br />

AM<br />

AM–PB<br />

NT<br />

AM<br />

NT<br />

NT<br />

Colomb<br />

NT<br />

TRAn<br />

And<br />

Mexico<br />

Mexico<br />

Argent<br />

AM<br />

AM<br />

Andes<br />

AM–PE<br />

Bah,Cuba<br />

SAm<br />

BR–SM<br />

NT<br />

NT<br />

AM<br />

NT<br />

NT<br />

NT<br />

CBR<br />

SAm<br />

SAm<br />

SBR<br />

Argent<br />

50<br />

Mim d Myr e<br />

– –?<br />

–<br />

–<br />

– –?<br />

+ –<br />

–<br />

– –?<br />

– –?<br />

+– +–<br />

–<br />

–?<br />

– –?<br />

– –?<br />

– –?<br />

–<br />

– –?<br />

– –?<br />

– –?<br />

–?<br />

– –?<br />

– –?<br />

–<br />

+?<br />

– +<br />

– +<br />

+ + +?<br />

+?<br />

+– +<br />

+– +<br />

+ ?<br />

+?<br />

+– +<br />

+ ?<br />

+ ?<br />

Time<br />

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

activity<br />

PM<br />

N,AM<br />

AM<br />

?<br />

AM<br />

PM<br />

?<br />

MD<br />

MD<br />

PM<br />

PM<br />

PM<br />

?<br />

9<br />

MD<br />

9<br />

?<br />

PM<br />

PM<br />

?<br />

AM,PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

MD<br />

MD<br />

MD<br />

MD<br />

Habitat<br />

(usual) c<br />

F.Cd.Cmp<br />

HumidF<br />

HumidF<br />

RainF<br />

SwampF<br />

HumidF<br />

?<br />

HumidF<br />

F,Cd<br />

RainF<br />

HumidF<br />

HumidF<br />

CloudF<br />

HumidF<br />

Cd,Cmp<br />

?<br />

?<br />

HumidF<br />

CloudF<br />

?<br />

Coast<br />

Scrub<br />

HumidF<br />

CloudF<br />

Cd,Cmp<br />

F,Cd<br />

HumidF<br />

HumidF<br />

HumidF<br />

HumidF<br />

Cd,Cmp<br />

Cd,Cmp<br />

Cd,Cmp<br />

Cd,Cmp<br />

Chaco<br />

Abn f<br />

4<br />

4<br />

2<br />

?<br />

4<br />

2<br />

?<br />

3<br />

4<br />

3<br />

2<br />

2<br />

2<br />

?<br />

3<br />

?<br />

?<br />

?<br />

3<br />

?<br />

3<br />

2<br />

2<br />

4<br />

3<br />

2<br />

2<br />

3<br />

3<br />

4<br />

3<br />

4<br />

2<br />

2<br />

Larval<br />

host<br />

plants q<br />

?<br />

Epiphylls<br />

Euphorbi.<br />

?<br />

Marant., Ar.<br />

Clusi.<br />

?<br />

Legumin.,<br />

Myrt.<br />

?<br />

?<br />

?<br />

?<br />

?<br />

Ros., Legumin.<br />

?<br />

?<br />

?<br />

?<br />

?<br />

?<br />

?<br />

?<br />

Euphorbi.<br />

Legumin.,<br />

Euphorbi.<br />

?<br />

?<br />

Legumin.,<br />

Euphorbi.<br />

Simaroub.,<br />

Legumin.,<br />

Sterculi.,<br />

Euphorbi.<br />

Legumin.<br />

?<br />

Legumin., Ros.,<br />

Turner.<br />

?<br />

?<br />

Bibliography<br />

3<br />

2,3,23<br />

17<br />

17<br />

17<br />

2,3<br />

10<br />

21<br />

20<br />

11<br />

18<br />

12<br />

23<br />

14<br />

19


Table 1 (cont). Synopsis <strong>of</strong> neotropical Riodininae a ; systematics and biology 1 *.<br />

Taxonomic groups:<br />

SUBFAMILY (TRIBE),<br />

Subtribe or Group,<br />

* † Genus;<br />

Nymphidiiti<br />

Parnes<br />

Periplacis<br />

†<br />

Menander<br />

*Zelolaea<br />

Pandemos<br />

† *Dysmalhia<br />

Joiceya<br />

Rodinia<br />

Calociasma<br />

† *Calospila<br />

† *Adeloiypa<br />

*† Selabis (=Orimba)<br />

† *Theope<br />

† *Nymphidium<br />

Stalachtiti<br />

Stalachtis<br />

Plate<br />

numbers<br />

(Lewis<br />

1973)<br />

77:21<br />

77:23<br />

75:4,7–9<br />

78:9<br />

79:39<br />

77:19<br />

72:6<br />

–<br />

78:16<br />

71:6<br />

71:3–4<br />

:21,74<br />

77:27<br />

:30–5<br />

78:1–6<br />

72:3,7–12<br />

:15–6<br />

77:5–12<br />

78:20,33<br />

79:11–26<br />

76:18<br />

:20–5<br />

78:29–30<br />

:32,35<br />

Approx.<br />

no. <strong>of</strong><br />

spp.<br />

2<br />

1<br />

8<br />

2<br />

2<br />

7<br />

1<br />

1<br />

4<br />

40<br />

31<br />

29<br />

45<br />

34<br />

8<br />

Typical or<br />

well-known<br />

species<br />

philotes<br />

glaucoma<br />

hebrus<br />

phasma<br />

pasiphae<br />

portia<br />

praeclara<br />

calpharnia<br />

lilina<br />

emylius,<br />

zeanger<br />

senta, bolena<br />

epitus, lagus,<br />

cruentata<br />

terambus<br />

lisimon<br />

phaedusa<br />

Distribution<br />

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

genus c<br />

AM–Atl<br />

SAm<br />

NT<br />

BR<br />

NT<br />

AM<br />

BR–MT<br />

AM<br />

NT<br />

NT<br />

NT<br />

NT<br />

NT<br />

NT<br />

SAm<br />

Notes:<br />

* Genus in which 'species' names will probably be united, reducing total<br />

number <strong>of</strong> species.<br />

† Genus in which new species will be described, increasing total number <strong>of</strong><br />

species (If both symbols present, the first one should predominate -<br />

overall increase or reduce).<br />

a Does not include primarily Nearctic species extending into Mexico.<br />

b Information on Riodininae from Callaghan and Lamas (1994) and Harvey<br />

(1987)<br />

c AM = Amazon; And = Andes; Ant = Antilles; Argent = Argentina; Atl =<br />

Atlantic; BA = Bahia; Bah = Bahamas; BR = Brazil; CBR = Central<br />

Brazil; Cd = Cerrado; C\ = Cloud; Colomb = Colombia; Cmp = Campo;<br />

CR = Costa Rica; EcAnd = Ecuadorian Andes; F = Forest; Hu = Humid;<br />

MG = Minas Gerais; MT = Mato Grosso, Brazil; NA = North America;<br />

NT = Neotropics; PE = Pernambuco, NE Brazil; SAm = South America;<br />

SBR = Southern Brazil; Sec = Secondary forests; SM = Serra do Mar;<br />

sp.AM = especially the Amazon; TRAn = TransAndean; upAM = Upper<br />

Amazon basin; Venez = Venezuela.<br />

d Mim = mimicry <strong>of</strong> distasteful or venomous species; '++' = strongly<br />

present or characteristic,' +' = present,' +-' = weak, variable or sometimes<br />

present.<br />

e Myr = larva myrmecophilous.<br />

f Abn = usual abundance when found: 1 = very scarce, 2 • scarce, 3 =<br />

uncommon, 4 = common, 5 = abundant.<br />

g Maintained at family level by Harvey and others; subfamilies = Styg,<br />

Corrach, Hamear, Eus, Rio.<br />

h Divided by Harvey into Mesosemiini (first 5 genera), Eurybiini (nos<br />

7,8,9), 'incertae sedis' (Eunogyra and rest).<br />

51<br />

Mini d<br />

–<br />

–<br />

–<br />

+<br />

–<br />

–<br />

–<br />

–<br />

–<br />

+–<br />

–<br />

++<br />

+–<br />

+–<br />

++<br />

Myr e<br />

+?<br />

+?<br />

+<br />

+<br />

+?<br />

+?<br />

+?<br />

+?<br />

+?<br />

+<br />

+<br />

+?<br />

+<br />

+<br />

+<br />

Time<br />

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

activity<br />

PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

?<br />

?<br />

PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

PM<br />

MD<br />

Habitat<br />

(usual) c<br />

RainF<br />

HumidF<br />

HumidF<br />

RainF<br />

HumidF<br />

RainF<br />

Cd<br />

?<br />

OpenF<br />

HumidF<br />

HumidF<br />

RainF<br />

F,Cd<br />

HumidF<br />

F,Cd,Cmp<br />

Abn f<br />

2<br />

2<br />

2<br />

1<br />

1<br />

1<br />

1<br />

1<br />

2<br />

4<br />

4<br />

2<br />

2<br />

4<br />

4<br />

Larval<br />

host<br />

plants')<br />

?<br />

?<br />

Marcgravi.<br />

?<br />

?<br />

?<br />

?<br />

?<br />

?<br />

Malpighi.<br />

?<br />

Slerculi. o<br />

Legumin.,<br />

Sterculi.<br />

Legumin.,<br />

Sterculi.<br />

Simaroub.,<br />

Legumin.<br />

Bibliography<br />

15<br />

16<br />

22<br />

5<br />

23,24<br />

i Divided by Harvey into 'Ancylurissection' (first 17 genera) and 'Riodina<br />

section' (remaining 23).<br />

j Probably part <strong>of</strong> Rhetus, isolated in high mesic rockfields (see Figure 3<br />

and species account).<br />

k United into a single species by Azzará (1973); probably 3 species at most.<br />

1 Divided by Harvey into Sarotini and Helicopini (this tribe monotypic).<br />

m Callaghan maintains 20 species, but 4 (Harvey) is more likely.<br />

n Divided by Harvey into a'Lemonias-secuon' (first 3 genera), a 'Synargissection'<br />

(next 4), and an'Audre-section' (last 5 genera).<br />

0 Larvae have been reported feeding on Membracidae (carnivorous), the<br />

only case <strong>of</strong> aphytophagy known in the neotropical <strong>Lycaenidae</strong> to date.<br />

(Harvey, pers. comm., indicates probable aphytophagy in Mimocastnia).<br />

p A single fossil genus (near Rhetus) and species have been described:<br />

Riodinella nympha Durden and Rose 1978, from the middle Eocene,<br />

about 48 million years ago.<br />

q Plant family names abbreviated by omission <strong>of</strong> standard ending.<br />

Bibliography cited in right-hand column:<br />

1 Harvey 1989; 2 Downey and Allyn 1980; 3 Kendall 1976; 4 Horvitz et al. 1987;<br />

5 Callaghan 1989; 6 Azzara 1978; 7 Clench 1967; 8 McAlpine 1971;<br />

9 Clench 1972; 10 Harvey and Clench 1980; 11 Ross 1964, 1966;<br />

12 Callaghan 1982b; 13 Callaghan 1986b; 14 Schremmer 1978; 15 Callaghan<br />

1977, 1978; 16 Harvey and Gilbert 1988; 17 Callaghan 1982b; 18 De Vries<br />

1988b, 1991b; 19 Miller and Miller 1972; 20 Callaghan 1979; 21 Callaghan<br />

1983b; 22 Talbot 1928; 23 Dias 1980; 24 Callaghan 1986a.


Table 2. Synopsis <strong>of</strong> systematics and biology <strong>of</strong> neotropical coppers and blues.<br />

SUBFAMILY (TRIBE)<br />

Section<br />

Genus<br />

LYCAENINAE<br />

Heliophorus section<br />

lophanus<br />

Plate<br />

nos<br />

(Lewis)<br />

62:34 e<br />

POLYOMMATINAE (POLYOMMATINI) b<br />

Leplotes section<br />

Leptotes<br />

Zizula section<br />

Zizula<br />

Brephidium section<br />

Brephidium<br />

Everes section<br />

Everes<br />

'Everes' c<br />

Lycaenopsis section<br />

Celastrina<br />

Polyommatus section<br />

Hemiargus<br />

(+2 subgenera)<br />

Pseudochrysops<br />

Madeleinea d (Itylos auctt.)<br />

Paralycaeides d<br />

Pseudolucicia d<br />

Nabokovia d<br />

Itylos d (Parachilades auctt.)<br />

Polytheclus d<br />

(19:34)<br />

69:51<br />

(19:17)<br />

(19:13)<br />

–<br />

(19:11)<br />

(19:23)<br />

67:18<br />

–<br />

–<br />

–<br />

67:25–6<br />

–<br />

67:23<br />

–<br />

No.<br />

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

spp.<br />

1<br />

8<br />

2<br />

1<br />

1<br />

2<br />

1<br />

9<br />

1<br />

>8<br />

4<br />

19<br />

2<br />

4<br />

2<br />

Typical<br />

species<br />

pyrrhias<br />

cassius<br />

cyna<br />

exilis<br />

comyntas<br />

cogina<br />

ladon<br />

hanno<br />

bornoi<br />

pelorias<br />

inconspicua<br />

chilensis<br />

faga<br />

titicaca<br />

sylphis<br />

Distribution*<br />

CAm<br />

NT<br />

NT<br />

CAm–Ven<br />

CAm<br />

BR–SM<br />

CAm<br />

NT<br />

Ant<br />

SAnd<br />

SAnd<br />

SAnd<br />

SAnd<br />

SAnd<br />

SAnd<br />

Habitat a<br />

CloudF<br />

Cmp, Sec<br />

Swamp<br />

Estuary<br />

Cmp<br />

Cmp<br />

Cmp, Forest<br />

Cmp, Scrub<br />

Scrub<br />

Scrub<br />

Scrub<br />

Puna<br />

Scrub<br />

Scrub, Puna<br />

Scrub<br />

Abundance*<br />

2<br />

5<br />

3<br />

4<br />

3<br />

3<br />

4<br />

4<br />

2<br />

2<br />

2<br />

3<br />

2<br />

2<br />

2<br />

Larval Hosts<br />

?<br />

Legumin.<br />

Legumin.?<br />

Various<br />

Legumin.<br />

Legumin.?<br />

Legumin.<br />

Legumin.<br />

Notes<br />

a The conventions and abbreviations in these categories follow those in Table 1.<br />

b Does not include North American (Nearctic) species invading northern Mexico.<br />

c Dr. Heinz Ebert gave a new name to this genus, still unpublished; may be true Everes. The species griqua Schaus was renamed as Pseudolucia parana by<br />

Hálint (1993), but Dr. Ebert regarded this taxon as belonging to the Everes section and congeneric with cogina.<br />

d Following Bálint 1993.<br />

e This figure number from de la Maza 1988.<br />

General references for neotropical Polyommatinae include Nabokov 1945; Clench 1964; and the very useful catalogue <strong>of</strong> Bridges 1988.<br />

Notes to Table 3 (opposite).<br />

a Table composed with the help <strong>of</strong> Dr. R.K. Robbins. A number <strong>of</strong> well-known genera <strong>of</strong> Eumaeini are either little-known biologically, or with infrageneric<br />

relationships still poorly defined, and thus are not included in this Table. Genera briefly described by Johnson (1991 a) are also omitted, for lack <strong>of</strong> complete<br />

information on their scope and position; the same applies to the 20 new neotropical genera and 88 new species proposed by Johnson (1992) 'based on adult<br />

wing pattern, tergal morphology and male and female genitalia' (all <strong>of</strong> these show great variation in some populations), and to the two new genera and 29<br />

new species added by Johnson and coauthors, in numbers 23 (1992, with five separate papers) and 24–29 (1993, including a review <strong>of</strong> Pseudolucia and<br />

two new genera <strong>of</strong> Polyommatinae, see Table 2. The well known genera not included here include: Theritas, Mithras, Allosmaitia, Thereus (=Noreena),<br />

Ocaria, Parrhasius, Michaelus, Oenomaus, Symbiopsis (Nicolay 1971b), Calycopis s.l. (=Calystryma, Femniterga, Tergissima and 15 <strong>of</strong> the other genera<br />

described in Johnson 1991a; the other seven new 'Outgroup' genera also include several segregates from well-known genera mentioned here),<br />

Electrostrymon, Lamprospilus, Theclopsis, Siderus, Contrafacia (=Orcya), Ipidecla, Hypostrymon, Nesiostrymon (see Johnson 1991b, including its<br />

citations <strong>of</strong> most previous papers <strong>of</strong> that author), Paiwarria and Theorema.<br />

b The conventions and abbreviations in these categories follow those in Table 1.<br />

c References: 1 Robbins 1987; 2 Nicolay 1971a; 3 Robbins 1980,1985; 4 Robbins 1991; 5 Nicolay 1980 and Johnson 1989; 6 Nicolay 1977; 7 Clench 1944,<br />

1946; 8 Nicolay 1980 and Callaghan 1982; 9 Nicolay 1982; 10 Miller 1980; 11 Robbins and Venables 1991.<br />

52


Table 3. Synopsis <strong>of</strong> the systentatics and biology <strong>of</strong> selected neotropical hairstrcaks (better defined genera, including about 30% <strong>of</strong> known species) a .<br />

SUBFAMILY (TRIBE),<br />

Genus<br />

Atlides<br />

Areas<br />

Pseudolycaena<br />

Arawacus (=Polyniphes,<br />

Dolymorpha)<br />

Rekoa (=Heterosmailia)<br />

Chlorostrymon<br />

Magnastigma<br />

Cyanophrys<br />

Panthiades (=Cycnus)<br />

Olynthus<br />

Strymon<br />

Tmolus<br />

Ministrymon<br />

Brangas<br />

Chalybs<br />

Erora<br />

Trichonis<br />

laspis<br />

Janlhecla<br />

Plate<br />

numbers<br />

(Lewis)<br />

THECLINAE (EUMAEINI)<br />

Eumaeus<br />

67:17<br />

Theslius<br />

69:7<br />

Micandra<br />

67:47,51<br />

69:8<br />

Evenus<br />

69:13,19<br />

:25,45<br />

66:12<br />

:14–6<br />

68:8<br />

69:44,46<br />

68:39<br />

66:9–11<br />

67:24<br />

68:26<br />

69:17,21<br />

67:45<br />

68:12,44,48<br />

69:9<br />

67:30–1<br />

67:29,36<br />

68:15<br />

:23–4<br />

:34<br />

67:22,44<br />

68:5<br />

–<br />

66:18<br />

67:1,3<br />

:28<br />

68:19,53<br />

69:40<br />

:47–8<br />

69:49,52<br />

67:2,15<br />

:27<br />

68:41<br />

69:38<br />

66:13<br />

68:27<br />

68:20<br />

–<br />

69:18,27<br />

:41<br />

68:38<br />

No.<br />

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

spp.<br />

6–7<br />

1–3<br />

8–10<br />

13<br />

13<br />

7<br />

4–6<br />

16–20<br />

7<br />

5<br />

6<br />

19–22<br />

8<br />

10<br />

40–70<br />

10–11<br />

18–22<br />

18–20<br />

3<br />

>40<br />

2<br />

11–12<br />

10<br />

Typical<br />

species<br />

minijas<br />

pholeus<br />

platyptera<br />

regalis<br />

polybe<br />

imperialis<br />

marsyas<br />

aetolus<br />

dumenilii<br />

melon<br />

marius<br />

palegon<br />

simaethis<br />

julia, hirsuta<br />

herodotus<br />

bertha<br />

acaste<br />

bitias, phaleros<br />

punctum, fancia<br />

ziba<br />

mulucha<br />

echion<br />

azia<br />

una<br />

getus, silumena<br />

janias<br />

aura, phrosine<br />

hyacinthinus<br />

temesa, talayra<br />

janthina, rocena<br />

Distribution<br />

b<br />

Ant–AM<br />

AM<br />

CAm–And<br />

NT<br />

NT<br />

NT<br />

NT<br />

NT<br />

NT<br />

NT<br />

NT<br />

NT<br />

NT<br />

CR–SBR<br />

NT<br />

NT<br />

NT<br />

NT<br />

NT<br />

NT<br />

AM<br />

NT<br />

NT<br />

53<br />

Habitat b<br />

F, Scrub<br />

HumidF<br />

CloudF<br />

Cl, HumidF<br />

Cl, Humid<br />

RainF<br />

Cl, Humid<br />

RainF<br />

F, Cd, Scrub<br />

F, Cd, Scrub<br />

F, Cd, Cmp<br />

F, Cd, Cmp<br />

Hu, RainF, Cd<br />

HumidF<br />

Cl, Hu, RainF<br />

Cl, RainF<br />

F, Cd, Cmp<br />

Hu, RainF, Cd<br />

HumidF, Scrub<br />

Cl, Hu, RainF<br />

HumidF<br />

Cl, Hu, RainF<br />

RainF<br />

Forest<br />

Cl, Hu, RainF<br />

Abundance<br />

3<br />

2<br />

2<br />

2<br />

2<br />

3<br />

4<br />

4<br />

4<br />

2<br />

2<br />

3<br />

3<br />

2<br />

4<br />

3<br />

2<br />

2<br />

2<br />

2<br />

1<br />

2<br />

3<br />

Larval<br />

hosts<br />

Cycad.<br />

Legumin.<br />

Sapot.<br />

Loranth.<br />

Laur.,<br />

Anacardi.<br />

Polyphagous<br />

Solan.<br />

Polyphagous<br />

Sapind.<br />

Polyphagous<br />

Polyphagous<br />

Lecythid.<br />

Polyphagous<br />

Polyphagous<br />

Legumin.<br />

Myr. b<br />

?<br />

–<br />

–<br />

–<br />

–<br />

–<br />

–<br />

+<br />

+<br />

+<br />

+?<br />

+<br />

+<br />

+<br />

+<br />

Loranth., Sapind.<br />

Legumin.<br />

Polyphagous<br />

?<br />

Sterculi.<br />

?<br />

+<br />

Refs c<br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

1<br />

11


Figure 2. Biogeographical patterns <strong>of</strong> neotropical <strong>Lycaenidae</strong> and other butterflies.<br />

54


On the basis <strong>of</strong> four foreleg characters, Robbins (1988) has<br />

advocated separation <strong>of</strong> the Riodininae from the <strong>Lycaenidae</strong>,<br />

suggesting affinity with the Nymphalidae. The considerable<br />

variation <strong>of</strong> some leg characters leads to difficulty and<br />

disagreement about their interpretation in many groups and in<br />

this case further integration with other sets <strong>of</strong> characters would<br />

be helpful before a final decision is reached.<br />

Distribution and variation<br />

As noted above, three species <strong>of</strong> neotropical blues (Hemiargus<br />

hanno, Leptotes cassias and Zizula cyna) are very widespread<br />

indeed despite their minute size and weak flight, especially<br />

Zizula cyna which struggles to rise above the grasses in the sites<br />

it inhabits throughout tropical America. Leptotes and Hemiargus<br />

are also known from distant oceanic islands (Galapagos and<br />

Fernando de Noronha), as well as all islands in the Caribbean.<br />

Robbins and Small (1981) have shown that most Theclinae are<br />

also widespread and undifferentiated, and suggest that this is<br />

due to mass, long-range, community dispersal during the dry<br />

season, like Pieridae and Hesperiidae, ranging over the entire<br />

region. Riodininae may be at the opposite extreme <strong>of</strong> the<br />

spectrum, most being extremely local (even in more widespread<br />

species) and <strong>of</strong>ten well-differentiated geographically (especially<br />

in Eurybia, Mesosemia, Nymphidium and the mimetic groups).<br />

Largely endemic thecline faunas are found in high mountains<br />

in southeastern Brazil, the Andean cloud forests and the Central<br />

American highlands, while local riodinine assemblages inhabit<br />

wet lowland forests (Callaghan 1983a, 1985). A largely endemic<br />

fauna (Table 2) inhabits the extremely long, narrow 'island'<br />

forming the Republic <strong>of</strong> Chile, isolated by desert (N), highaltitude<br />

rocks (E), ice (S) and ocean (W). There are contiguous<br />

areas in southwestern Bolivia and Peru extending up the semiarid<br />

'puna' as far as southern Ecuador in some cases. In the<br />

Brazilian Atlantic region (Figure 2) much endemism is seen in<br />

cerrado species (central plateau) in addition to montane and<br />

coastal forms; in the overall region, 49% <strong>of</strong> the Theclinae, 44%<br />

<strong>of</strong> the Riodininae, and two congeneric blues are endemic. These<br />

levels are slightly higher than for skippers, nymphalids or<br />

pierids (all 39%) or papilionids (42%), but they may drop when<br />

proper links can be discovered between Atlantic and Amazonian<br />

or Andean sister-groups (species or subspecies).<br />

Robbins (pers. comm.) has suggested a list <strong>of</strong> 25 thecline<br />

species or superspecies which are 'widespread, common in<br />

many regions and habitats, and likely to be recorded on a quick<br />

trip to the neotropics': Theritas mavors/triquetra, Pseudolycaena<br />

supersp. marsyas, Arawacus supersp. aetolus, Rekoa meton, R.<br />

marius, R. palegon, Tmolus echion, Oenomaus ortygnus,<br />

Panthiades bitiaslhebraeus, Parrhasiuspolibetes, Cyanophrys<br />

herodotus, Electrostrymon supersp. endymion, Calycopis<br />

isobeon, C. cerata, Chlorostrymon simaethis, Ministrymon<br />

una, M. azia, Strymon mulucha, S. bazochii, S. ziba, and<br />

'Thecla' (catch-all genus for the majority <strong>of</strong> species still not<br />

assigned to existing generic names) hemon, hesperitis, syllis,<br />

55<br />

celmus and tephraeus. In South America, Arcas imperialis,<br />

Panthiades phaleros and some other forest species could be<br />

added to this list.<br />

At the other extreme, one or two specimens are known from<br />

a single locality (such as two new Atlides species from a single<br />

2200m hilltop west <strong>of</strong> Cali, Colombia), and perhaps over a<br />

hundred (about 10%) are known from fewer than ten specimens<br />

in museum collections. In April 1991, Robbins took the fourth<br />

known specimen <strong>of</strong> a striking thecline, without generic or<br />

specific identification, on Mikania flowers invadingthis author's<br />

backyard; the first three were in a single Brazilian collection,<br />

awaiting description. So for these species, and perhaps for over<br />

50% <strong>of</strong> Riodininae, much information is still waiting in the<br />

wings before any idea <strong>of</strong> distribution, biology, affinities or<br />

importance can be assigned. The general patterns <strong>of</strong><br />

biogeography <strong>of</strong> neotropical <strong>Lycaenidae</strong> are very similar to<br />

those known in other insects, as shown in Figure 2.<br />

Mimetic species that join 'rings' <strong>of</strong> similar and distasteful<br />

heliconians, ithomiines, arctiids, dioptids and other Lepidoptera<br />

vary in parallel with these (mostly in 16 genera <strong>of</strong> Riodininae,<br />

Table 1) and are just as useful as their models for identifying<br />

centres <strong>of</strong> endemism at the infraspecific level in the lowland<br />

neotropical forests (Figure 2). Many other <strong>Lycaenidae</strong> show<br />

strong geographic variation, <strong>of</strong>ten regarded today as at the<br />

species rather than the subspecies level. Probably about half <strong>of</strong><br />

any local lycaenid fauna will contain some information about<br />

region and habitats.<br />

Behaviour and juvenile host plant<br />

relations<br />

Most neotropical blues and hairstreaks have the usual rapid,<br />

darting flight <strong>of</strong> these groups. Especially impressive are the<br />

larger hilltopping species <strong>of</strong> Arcas, Evenus, Atlides and Brangas,<br />

but they are surely no faster than some <strong>of</strong> the small species<br />

which are completely lost by the eye on each flight and are only<br />

seen again when they land back on the same perch.<br />

Riodinine flight varies from the same rapid, darting pattern<br />

(Theope, many Euselasia, Charis, Mesene, Symmachia,<br />

Stichelia, Calydna, Calospila) through a more usual and slower,<br />

erratic flight (always returning to a chosen perch) to the very<br />

casual, dipping or fluttering flight <strong>of</strong> mimetic species imitating<br />

their distasteful models (Methone, Ithomiola, Ithomeis,<br />

Brachyglenis, Themone, Chamaelimnas, Cartea, Uraneis,<br />

Stalachtis, and females <strong>of</strong> Esthemopsis and Setabis). A few<br />

species may imitate wasps with a 'buzzy' flight (Chorinea,<br />

Syrmatia, some Rhetus), while ground-perching species may<br />

rise in spirals when disturbed. Styx has a very weak, almost<br />

falling flight (fide G. Lamas).<br />

Adults usually visit small flowers, and many riodinines<br />

(principally males) are also found on damp sand or mud<br />

(Lyropteryx, Riodina, Rhetus, Necyria, Barbicornis, Parcella,<br />

Notheme, Monethe, Lasaia, Chamaelimnas, Caria, Siseme).<br />

All vertical levels <strong>of</strong> the habitat, and also adjacent non-forest


areas, are included in usual foraging surveys.<br />

Adults <strong>of</strong> the genus Eumaeus, locally common from southern<br />

Florida to central-west Brazil, incorporate toxic cycasins from<br />

their larval foodplants (mostly Zamia but <strong>of</strong>ten other<br />

Cycadaceae) (Rothschild et al. 1986; Bowers and Larin 1989).<br />

They are brightly marked models for mimicry rings, flying very<br />

slowly and liberating cyanide gas when crushed; their larvae<br />

are gregarious and also highly aposematic, having a delicate<br />

relationship with the phenology <strong>of</strong> their hosts (Clark and Clark<br />

1991). No other adult neotropical <strong>Lycaenidae</strong> have yet been<br />

shown to be similarly protected against predators.<br />

Mass movements <strong>of</strong> Riodininae communities have not been<br />

reported, though populations <strong>of</strong>ten appear suddenly in a new<br />

habitat as a few worn individuals and are later replaced by a<br />

large endogenous generation if foodplants are found. Theclinae<br />

seem to move over large regions in loose, multispecific groups<br />

(hundreds <strong>of</strong> individuals <strong>of</strong> dozens <strong>of</strong> species), or sometimes<br />

migrate unidirectionally or erratically as dense populations <strong>of</strong><br />

a single species (Pseudolycaena and Calycopis), principally<br />

during the dry season when flowers are scarce. Sunny clearings<br />

can attract such concentrations, which can at times contain<br />

nearly a hundred species. Especially interesting areas to find<br />

these assemblages are sheltered sites on low or high ridges, on<br />

dry, sunny or windy days (Robbins and Small 1981).<br />

Territorial perching and area defence by males are widespread<br />

and easily demonstrated by marking neotropical <strong>Lycaenidae</strong><br />

(see also Bates 1859). In open fields, males perch on the tips <strong>of</strong><br />

the highest grass blades, driving <strong>of</strong>f with their rapid forays other<br />

males, many other butterflies and even insect predators, bird<br />

predators and collectors. In forests, perching may be in the<br />

canopy (only in the morning since canopy species descend to<br />

the cooler understorey at noontime), on forest edges, in small<br />

clearings, on tree trunks in the sun or shade, on top <strong>of</strong> or under<br />

leaves at all levels (with distinct partitioning, Callaghan 1983a),<br />

along rivercourses or trails, or on hilltops and rocks (especially<br />

in mountains), almost always in an exposed, <strong>of</strong>ten sunlit spot.<br />

Territories defended vary from less than one to many dozens <strong>of</strong><br />

cubic metres. As might be expected, hilltopping (or ridgetopping<br />

or edge-seeking) is very common in male <strong>Lycaenidae</strong><br />

in wait for females <strong>of</strong> their sparse populations. More than one<br />

hundred species may accumulate at a favoured hilltop in late<br />

morning to mid-afternoon, colouring the sky in their territorial<br />

battles. Early morning 'leks' <strong>of</strong> many males have been seen in<br />

Euselasia, Barbicornis, Sarota and Syrmatia.<br />

When not exhibiting territorial behaviour, theclines perch<br />

on the tops <strong>of</strong> leaves or on twigs in the forest undergrowth, in<br />

shade or sun-flecks, 'rubbing' their hindwings in the usual<br />

manner to call attention to their 'false head' (Figure 3 and<br />

Robbins 1980,1985) (this also occurs in the nodminzs Anteros<br />

and Sarota). Whilst doing this, they fly upon the slightest<br />

provocation, usually landing quite far away.<br />

Riodinines usually perch under leaves with wings flat open<br />

(most genera) or closed (Euselasiini, Sarota, Anteros, Themone,<br />

Helicopis, Theope, most Setabis); some perch on tops <strong>of</strong> leaves<br />

with wings half-open (Mesosemia, Semomesia and relatives,<br />

also Eurybia at dusk). Particular underleaf perches are used<br />

56<br />

repeatedly and insistently by the same individual, and by<br />

members <strong>of</strong> the same population. If the resident individual is<br />

removed from a leaftop territorial perch or an underleaf resting<br />

station, a series <strong>of</strong> conspecific males will then occupy the same<br />

perch; depending upon the genus, these can be progressively<br />

younger and more splendid, or older and more worn than the<br />

first resident.<br />

Courtship can have both aerial and perched components,<br />

either <strong>of</strong> which can be very complex. They are presumably<br />

accompanied by diverse pheromonal signals and responses<br />

which are produced by scent-spots or pads, androconial brushes<br />

or patches. In Helicopis cupido (observed in Guyana) the male<br />

alternates rapid wing flutters (to 40° open) with short 220° wing<br />

opened phases (one-second cycles), whilst perched right behind<br />

the female (who maintains her wings closed) on top <strong>of</strong> a sloping<br />

large leaf <strong>of</strong> the larval foodplant (Montrichardia, Araceae); he<br />

then moves directly in to initiate mating.<br />

Oviposition is not <strong>of</strong>ten observed, but as with most butterflies,<br />

occurs in short bouts which involve appreciable searching,<br />

inspection, and tactile evaluation (foreleg 'drumming') <strong>of</strong> the<br />

exact site. Off-hostplant egg-laying is not common but has been<br />

observed, as in other butterflies. Eggs are <strong>of</strong>ten placed in axils<br />

and other tight places, and are very rarely found; they are <strong>of</strong>ten<br />

pincushion-shaped and highly sculptured (Downey and Allyn<br />

1979,1980,1981,1984), but in four small tribes <strong>of</strong> Riodininae,<br />

they are smooth and barrel-shaped (Harvey 1987).<br />

Myrmecophilous species may lay their eggs along ant foraging<br />

routes (Stalachtis, Lemonias, Mycastor, Nymphidium) or even<br />

visually encourage ants to directly remove eggs as they are<br />

expelled (Adelotypa senta;Audre domina, Robbins and Aiello<br />

1982).<br />

Larvae may be smooth (especially if myrmecophilous),<br />

tubercled, spiny or hairy, <strong>of</strong>ten in relation to their habitat. They<br />

are usually highly cryptic, even when on flowers (a habit<br />

common in Theclinae, rare in Riodininae), and may be<br />

polymorphic, probably incorporating pigments from the flowers<br />

on which they feed (Monteiro 1991). Some larvae and many<br />

pupae are found in 'nests' inside rolled leaves. Pupae are quite<br />

variable in shape, with systematic significance in the Riodininae<br />

(Harvey 1987). These biological characters are quite similar to<br />

those <strong>of</strong> other <strong>Lycaenidae</strong>, though some riodinine larvae are<br />

unique.<br />

Myrmecophily is quite frequent in Theclinae, perhaps in<br />

50% <strong>of</strong> species (De Vries 1990; Malicky 1970), but in Riodininae<br />

has been reported in only three subtribes with 283 species:<br />

Eurybiiti; Lemoniiti; and Nymphidiiti (Harvey 1987; De Vries<br />

1991a). Stalachtis, in a fourth monogeneric subtribe, is optionally<br />

myrmecophilous, even in a single population (susanna, phlegia,<br />

lineosa, on Simaroubaceae; Callaghan 1986; Benson, Francini<br />

and Brown, unpublished).<br />

Ant-associated larvae have easily recognisable specialised<br />

glands (Cottrell 1984). In <strong>Lycaenidae</strong> these include a dorsal<br />

nectary organ on the seventh abdominal segment and a pair <strong>of</strong><br />

tentacle organs on the eighth segment. All known ant-associated<br />

riodinine larvae have a pair <strong>of</strong> glands (tentacle nectary organs)<br />

on the eighth segment. These secrete a solution which may be


57<br />

Figure 3. Pictures <strong>of</strong> Riodinines and Theclines in life, to show behaviour<br />

and habitat.<br />

(a) Paiwarria telemus (Manaus) [top left]; (b)Arawacus meliboeus (Japi, Sao<br />

Paulo) [top right]; (c) Strymon oreala (Santa Teresa, Espírito Santo) [middle<br />

left]; (d) Nirodia belphegor (Serra do Cip6, Minas Gerais) [middle right];<br />

(e) Helicopis acis (Belém, Para) [bottom]. (Photos: K.S. Brown Jr., except<br />

(d) Ivan Sazima.)


very rich in amino acids (De Vries 1988b; De Vries and Baker<br />

1989). Larvae <strong>of</strong> Lemoniiti and Nymphidiiti have two additional<br />

myrmecophilous organs (Harvey 1987): a pair <strong>of</strong> anterior<br />

tentacle organs on the metathorax which produce chemicals<br />

affecting ant behaviour (De Vries 1988b); and a pair <strong>of</strong> vibratory<br />

papillae on the anterior margin <strong>of</strong> the prothorax, which produce<br />

sound to call ants (De Vries 1988b, 1990), also seen in other<br />

lycaenids (De Vries 1991a, 1991b).<br />

Food resources <strong>of</strong> larvae are exceedingly varied, with at<br />

least 50 plant families recorded more than once; no pattern can<br />

be discerned except for certain tendencies for given species,<br />

genera or tribes to use the same resource over a wide geographical<br />

range. In contrast, some species are extremely polyphagous,<br />

especially those that feed on flowers (mostly Theclinae: Robbins<br />

and Aiello 1982; Monteiro 1991).<br />

The normally Myrtaceae-feeding and rare Euselasia eucerus<br />

in Brazil has taken to imported Eucalyptus leaves and becomes<br />

extremely abundant and destructive in commercial plantations<br />

<strong>of</strong> these Australian trees. The riodinine Audre campestris, the<br />

thecline Michaelus jebus, and some blues may be pests <strong>of</strong><br />

cultivated beans and other legumes, and Strymon ziba can<br />

become a serious pest on pineapples, but in general few<br />

<strong>Lycaenidae</strong> have achieved such notoriety.<br />

Lycaenids as indicator species<br />

The extremely sporadic distribution <strong>of</strong> most neotropical<br />

<strong>Lycaenidae</strong>, not only in collections but also in the field in both<br />

time and space, was noted by early naturalists (Bates 1859) and<br />

amply confirmed by all later observers. Riodinines are especially<br />

local, confined to a very narrow microhabitat, and active only<br />

at a particular time <strong>of</strong> day and level <strong>of</strong> forest; they are <strong>of</strong>ten<br />

greatly reduced in diversity and numbers in highly variable or<br />

unpredictable climates. Some are even nocturnal: Euselasia<br />

clesa was seen flying in typical territorial behaviour under a<br />

black light at 4 a.m., one hour before dawn and Sarota chrysus<br />

flies during the night and <strong>of</strong>ten comes to light (see also Miller<br />

1970). Most species, however, pick a time slot in the early<br />

morning or mid-afternoon to be active, with the latest species<br />

<strong>of</strong>ten being those <strong>of</strong> Nymphidium and Eurybia, the earliest<br />

Euselasia, Syrmatia and Sarota. While many species are present<br />

year round, great seasonal variation in both presence and<br />

abundance is the rule (as for Theclinae), with species <strong>of</strong>ten<br />

peaking either in mid-summer, late fall or late spring in higher<br />

to lower elevations. At least one species in perhumid tropical<br />

forest (Euselasia zara in southeastern Brazil) has been seen<br />

only in November (in lowlands) to February (in uplands) and<br />

seems to be univoltine, as has also been suggested for Audre<br />

domina in Panama (Robbins and Aiello 1982).<br />

All this natural fluctuation leads to serious problems in<br />

establishing baseline data for lycaenids, or recognising any<br />

significant tendencies to change or vary coherently in different<br />

habitats. When added to the great difficulty in collecting and<br />

identifying most species, the tendency to move about over the<br />

58<br />

landscape, and the chaotic state <strong>of</strong> the systematic and biological<br />

data, these characters greatly diminish the utility <strong>of</strong> neotropical<br />

<strong>Lycaenidae</strong> as ecological indicators at the present time.<br />

This does not mean that they are not potentially very useful<br />

in surveys and monitoring <strong>of</strong> natural and altered systems.<br />

Riodinines compose nearly half the daily list <strong>of</strong> butterflies in<br />

most parts <strong>of</strong> the Amazon Basin, and each species seems to be<br />

delicately tuned into a large number <strong>of</strong> physical and biological<br />

factors in its environment, which are therefore faithfully<br />

indicated by the species' presence (though not excluded by its<br />

absence). Some species may feed on a single plant genus or<br />

family as larvae or may be associated with certain ant species,<br />

thus necessitating the presence <strong>of</strong> these resources which may be<br />

harder to find and recognize than flying adult butterflies. The<br />

overall richness and diversity <strong>of</strong> the local lycaenid fauna <strong>of</strong>fers<br />

excellent opportunities for correlation with different systems in<br />

varying stages <strong>of</strong> naturalness. Even a partial list can lead to<br />

hypotheses about ecological factors, history <strong>of</strong> the region,<br />

resources present, soils and climate, primary productivity,<br />

importance <strong>of</strong> other communities, and stability <strong>of</strong> the whole<br />

system. The routine and very propitious use <strong>of</strong> <strong>Lycaenidae</strong> as<br />

indicators in the neotropics only awaits the resolution <strong>of</strong> the<br />

systematic picture in order to produce useable manuals for<br />

identification, and wider biological studies in order to expand<br />

the knowledge <strong>of</strong> indicator parameters possible with diverse<br />

species and groups.<br />

Threatened neotropical lycaenids<br />

Insects that are highly stenoecious – with many narrow<br />

environmental requirements or specialised interactions<br />

necessary for their survival – may be easily reduced or eliminated<br />

locally by minor habitat alteration through natural or human<br />

disturbance. This applies to many neotropical <strong>Lycaenidae</strong><br />

populations. With this local extinction, local adaptive genes<br />

will disappear, reducing the biodiversity and biosynthetic<br />

capacities <strong>of</strong> life.<br />

On the other hand, winged insects are <strong>of</strong>ten migratory and<br />

accustomed to seeking out the narrow environmental conditions<br />

needed for growth and reproduction, and are already adapted to<br />

normal levels <strong>of</strong> natural disturbance; they are likely to be<br />

widespread even though local, and very persistent in the regional<br />

species pool. Thus, a definition <strong>of</strong> 'threat' to a neotropical<br />

lycaenid should consider its ecological characteristics,<br />

geographical distribution, aptitude for mobility and colonisation,<br />

density <strong>of</strong> colonies, usual population size, voltinism (are adults<br />

always around to be able to flee destruction and colonise<br />

elsewhere?), and the kind, degree and extent <strong>of</strong> local or regional<br />

disturbance patterns – both natural (on long and short timescales)<br />

and artificial.<br />

In principle, essentially all lycaenid species should survive<br />

transformation <strong>of</strong> primitive mosaic habitats (in areas <strong>of</strong> complex<br />

topography) into anthropic mosaics with 10–30% <strong>of</strong> the original<br />

vegetation maintained in large patches. Such is the case in the


Brazilian Atlantic forests, where reduction <strong>of</strong> the original forest<br />

to 12% <strong>of</strong> its former area has not caused detectable extinction<br />

<strong>of</strong> lycaenids or indeed <strong>of</strong> any other insects or vertebrates<br />

monitored (Brown 1991; Brown and Brown 1991). In<br />

ecologically more homogeneous areas, reduction <strong>of</strong> natural<br />

habitat to 10% or less <strong>of</strong> its original area may still not lead to<br />

overall species extinction, though much genetic variation will<br />

be lost with local populations (and less vagile species in some<br />

groups can be eliminated). Most lycaenids will survive even in<br />

smallish habitats (0.1–100ha) as long as the edge effects (Lovejoy<br />

et al. 1986; Janzen 1984,1986) do not overwhelm the essential<br />

habitat characteristics: isolated areas <strong>of</strong> 10ha may be dramatically<br />

transformed within a year, and appreciable effects can be seen<br />

up to 250m from an edge in larger patches <strong>of</strong> experimental<br />

fragments in the central Amazon (Brown 1991). However,<br />

local disturbance can have a positive effect: it <strong>of</strong>ten brings in<br />

many new species <strong>of</strong> sun-loving lycaenids to the more diverse<br />

successional community (Hutchings 1991).<br />

In regions subjected to large-scale commercial conversion<br />

(to pasture, monoculture, silviculture, agrosilviculture or timber<br />

production), many lycaenids may not find new habitat and more<br />

large-scale regional extinction is possible. If a species is confined<br />

to the region, it might become extinct, especially if the habitat<br />

is very specialised and different from neighbouring areas (a<br />

basin, mountain top, headwater system, lakeshore or marsh,<br />

seasonally flooded region, or other natural 'island'). Thus,<br />

land-use patterns could make a large difference in the genetic<br />

erosion and threat to lycaenid species in tropical forests (Brown<br />

and Brown 1991).<br />

Brazil includes half the forest area, two <strong>of</strong> the four speciesendemic<br />

regions and 19 <strong>of</strong> the 45 subspecies-endemic centres<br />

in the neotropics; it has about 425 species <strong>of</strong> Theclinae and 700<br />

species <strong>of</strong> Riodininae (Brown 1982,1991). A new <strong>of</strong>ficial list<br />

<strong>of</strong> fauna threatened with extinction in Brazil was prepared by<br />

the Zoology Society in 1989 (Bernardes et al. 1990). It includes<br />

23 butterflies, only one <strong>of</strong> which is a lycaenid: Joiceya<br />

praeclarus, an inconspicuous monotypic genus known only<br />

from a small area in central Mato Grosso, and not seen there<br />

since its original discovery in the 1920s. This would represent<br />

the syndrome <strong>of</strong> 'restricted to a limited area, isolated, under<br />

intense large-scale human occupation', though it is probably<br />

preserved in the four conservation units established recently in<br />

the area. However, at least one population <strong>of</strong> the species must<br />

be found before its survival can be assured or even dealt with.<br />

A second list, prepared by the Zoology Society,<br />

recommending 'further study', includes another five species <strong>of</strong><br />

<strong>Lycaenidae</strong> and also six genera <strong>of</strong> little-known Riodininae<br />

which include mostly rare, extremely local and sometimes<br />

geographically limited species – the syndrome <strong>of</strong> 'widespread<br />

but extremely sporadically recorded' (Alesa, Colaciticus,<br />

Esthemopsis, Mesenopsis, Symmachia and Xenandra). To the<br />

generic list could be added Petrocerus, Erora, and other groups<br />

<strong>of</strong> riodinines and theclines confined to high-elevation habitat<br />

islands in southeastern Brazil (same syndrome as Joiceya, but<br />

somewhat more widespread).<br />

The five species <strong>of</strong> <strong>Lycaenidae</strong> on the second list represent<br />

59<br />

further syndromes <strong>of</strong> rarity and threat, which need some more<br />

study before inclusion on the <strong>of</strong>ficial Brazilian list <strong>of</strong> threatened<br />

fauna. They include the theclines Arcas ducalis (widespread<br />

but local in hills and mountains, mostly on hilltops) and Arawacus<br />

aethesa (highly restricted geographically to a lowland area<br />

under intensive deforestation); and the riodinines Eucorna<br />

sanarita (few localities in high mountains), Nirodia belphegor<br />

(a monotypic genus possibly part <strong>of</strong> Rhetus from natural<br />

rockfields in mountains <strong>of</strong> central Minas Gerais), and Helocopis<br />

cupido nr. lindeni (a few coastal swamps in the northeast).<br />

Other rare or restricted species may be added in the coming<br />

years to this list.<br />

Outside Brazil, the best candidates for threatened status may<br />

be island species or high-altitude groups, with most members<br />

still very little known. The primitive, taxonomically isolated<br />

Styx infernalis is but rarely seen at medium high elevations in<br />

Peruvian cloud forests. These and some further Brazilian cases<br />

are treated in the species accounts in this book.<br />

Acknowledgements<br />

Dr. Robert K. Robbins <strong>of</strong> the U.S. National Museum<br />

(Washington) kindly criticised the text <strong>of</strong> this chapter and<br />

produced the final version <strong>of</strong> Table 3, as well as providing<br />

extensive unpublished information on Theclinae. Drs Donald J.<br />

Harvey <strong>of</strong> the same institution, and Curtis John Callaghan <strong>of</strong><br />

Búzios, Rio de Janeiro and the Museu Nacional, furnished<br />

unpublished catalogues and extensive information on<br />

Riodininae. Reprints <strong>of</strong> papers and further notes on <strong>Lycaenidae</strong><br />

were provided by Kurt Johnson <strong>of</strong> the American Museum <strong>of</strong><br />

Natural History (New York) and Stan S. Nicolay <strong>of</strong> Virginia<br />

Beach. Dr. Olaf H.H. Mielke <strong>of</strong> the Departamento de Zoologia,<br />

Universidade Federal do Parana, Curitiba has helped greatly in<br />

knowledge <strong>of</strong> the Brazilian butterfly fauna over the past 30<br />

years, and curated a very rich and complete collection which<br />

has greatly aided in understanding systematics and<br />

biogeography.<br />

References<br />

AZZARÁ, M.L. 1978. Revisão do g6nero Barbicornis Godart, 1824<br />

(Lepidoptera, <strong>Lycaenidae</strong>, Riodininae). Acta Biol. Paran., Curitiba 7:<br />

23–69.<br />

BÁLINT, Z. 1993. A catalogue <strong>of</strong> the polyommatine <strong>Lycaenidae</strong> (Lepidoptera)<br />

<strong>of</strong> the xeromontane oreal biome in the Neotropics as represented in<br />

European collections. Repts Mus. Nat. Hist. Univ. Wisc. (Stevens Point)<br />

29: 1–36.<br />

BARCANT, M. 1970. <strong>Butterflies</strong> <strong>of</strong> Trinidad and Tobago. Collins, London,<br />

314pp.<br />

BATES, H.W. 1859. Notes on South American butterflies. Trans. ent. Soc.<br />

London N.S.5: 1–11.<br />

BERNARDES, A.T., MACHADO, A.B.M. and RYLANDS, A.B. 1990.<br />

Fauna Brasileira Ameacada de Extinçâo. FundacaoBiodiversitas/IBAMA,<br />

Belo Horizonte, 65pp.


BOWERS, M.D. and LARIN, Z. 1989. Acquired chemical defence in a<br />

lycaenid butterfly, Eumaeus atala. J. Chem. Ecol. 15: 1133–1146.<br />

BRIDGES, C.A. 1988. Catalogue <strong>of</strong> <strong>Lycaenidae</strong> and Riodinidae (Lepidoptera:<br />

Rhopalocera). Printed by the author, vii + 8 x ii + 888pp.<br />

BROWN JR., K.S. 1973. A Portfolio <strong>of</strong> Neotropical Lepidopterology. Rio de<br />

Janeiro, 28pp.<br />

BROWN JR., K.S. 1979 Ecologia Geográfica e Evoluqâo nas Florestas<br />

Neotropicais. Universidade Estadual de Campinas, xxxi + 265 + 120pp.<br />

BROWN JR., K.S. 1982. Historical and ecological factors in the biogeography<br />

<strong>of</strong> aposematic Neotropical butterflies. Amer. Zool. 22: 453–471.<br />

BROWN JR., K.S. 1991. <strong>Conservation</strong> <strong>of</strong> Neotropical palaeoenvironments:<br />

insects as indicators. In: Collins, N.M. and Thomas, J. A. (Eds), <strong>Conservation</strong><br />

<strong>of</strong> Insects and their Habitats. Academic Press, London, pp.349–404.<br />

BROWN JR., K.S. and BROWN, G.G. 1991. Habitat alteration and species<br />

loss in Brazilian forests: economic, social, biological and geological<br />

determinants. In: Whitmore, T.C. and Sayer, J.A. (Eds), Deforestation and<br />

Species Extinction in Tropical Forest, <strong>IUCN</strong>, Gland, Switzerland.<br />

CALLAGHAN, C.J. 1977. Studies on restinga butterflies. I. Life cycle and<br />

immature biology <strong>of</strong> Menander felsina (Riodininae), a myrmecophilous<br />

metalmark. J. Lepid. Soc. 31: 173–182.<br />

CALLAGHAN, C.J. 1978. Studies on restinga butterflies. II. Notes on the<br />

population structure <strong>of</strong> Menander felsina (Riodininae). J. Lepid. Soc. 32:<br />

87–94.<br />

CALLAGHAN, C.J. 1979. A new genus and a new subspecies <strong>of</strong> Riodinidae<br />

from Southern Brazil. Bull. Allyn Mus. 53: 1–7.<br />

CALLAGHAN, C.J. 1982a. Notes on immature biology <strong>of</strong> two myrmecophilous<br />

<strong>Lycaenidae</strong>: Juditha molpe (Riodininae) and Panthiades bitias<br />

(Lycaeninae).X Res. Lep. 20: 36–42.<br />

CALLAGHAN, C.J. 1982b. Three new genera <strong>of</strong> Riodininae from Mexico and<br />

Central America. Rev. Soc. Mex. Lep. 7: 55–63.<br />

CALLAGHAN, C.J. 1983a. A study <strong>of</strong> isolating mechanisms among<br />

Neotropical butterflies <strong>of</strong> the subfamily Riodininae. J. Res. Lep. 21:<br />

159–176.<br />

CALLAGHAN, C.J. 1983b. A new genus <strong>of</strong> Riodinine butterflies. Bull. Allyn<br />

Mus. 78: 1–7.<br />

CALLAGHAN, C.J. 1985. Notes on the zoogeographic distribution <strong>of</strong><br />

butterflies in the subfamily Riodininae in Colombia. In: Proc. 2nd Symp.<br />

on Neotropical Lepidoptera. J. Res. Lep., Supplement 1: 50–69.<br />

CALLAGHAN, C.J. 1986a. Notes <strong>of</strong> the biology <strong>of</strong> Stalachtis Susanna<br />

(<strong>Lycaenidae</strong>: Riodininae) with a discussion <strong>of</strong> Riodininae larval strategies.<br />

J. Res. Lep. 24: 258–263.<br />

CALLAGHAN, C.J. 1986b. Restinga butterflies: biology <strong>of</strong> Synargisbrennus<br />

(Stichel) (Riodinidae). J. Lepid. Soc. 40: 93–96.<br />

CALLAGHAN, C.J. 1989. Notes on the biology <strong>of</strong> three Riodinine species:<br />

Nymphidium lisimon attenuatum, Phaenochitonia sagaris satnius and<br />

Metacharis ptolemaeus(<strong>Lycaenidae</strong>: Riodininae). J. Res. Lep. 27:109–114.<br />

CALLAGHAN, C.J. and LAMAS, G. (in press). A checklist <strong>of</strong> the subfamily<br />

Riodininae. Atlas <strong>of</strong> Neotropical Lepidoptera.<br />

CLARK, D.B. and CLARK, D.A. 1991. Herbivores, herbivory and plant<br />

phenology: patterns and consequences in a tropical rain-forest cycad. In:<br />

Price, P.W., Lewinsohn, T.M., Fernandez, G.W. and Benson, W.W. (Eds),<br />

Plant-Animal Interactions: Evolutionary Ecology in Tropical and<br />

Temperate Regions, John Wiley and Sons, N.Y., pp. 209–225.<br />

CLENCH, H.K. 1944. Notes on Lycaenid butterflies, (a) The genus Callophrys<br />

in North America; (b) The acas/e-group <strong>of</strong> the genus Thecla. Bull. Harv.<br />

Mus. Comp. Zool. 94: 217–245.<br />

CLENCH, H.K. 1946. Notes on the amyntor group <strong>of</strong> the genus Thecla<br />

(Lepidoptera: <strong>Lycaenidae</strong>). Entomologist 79: 152–157, 185–191.<br />

CLENCH, H.K. 1964. A synopsis <strong>of</strong> the West Indian <strong>Lycaenidae</strong> with remarks<br />

on their zoogeography. J. Res. Lep. 2: 247–270.<br />

CLENCH, H.K. 1967. A note on Caria domitianus and ino (Riodinidae), with<br />

description <strong>of</strong> a new subspecies. J. Lepid. Soc. 21: 53–56.<br />

CLENCH, H.K. 1972. A review <strong>of</strong> the genus Lasaia (Riodinidae). J. Res. Lep.<br />

10: 149–180.<br />

COLLINS, N.M. and MORRIS, M.G. 1985. Threatened Swallowtail <strong>Butterflies</strong><br />

<strong>of</strong> the World: The <strong>IUCN</strong> Red Data Book. <strong>IUCN</strong>, Gland, 401pp.<br />

COTTRELL, C.B. 1984. Aphytophagy in butterflies: its relationship to<br />

60<br />

myrmecophily. Zool. J. Linn. Soc. 79: 1–57.<br />

DE VRIES, P.J. 1988a. The use <strong>of</strong> epiphylls as larval hostplants by the<br />

neotropical riodinid butterfly, Sarota gyas. J. nat. Hist. 22: 1447–1450.<br />

DE VRIES, P.J. 1988b. The larval ant-organs <strong>of</strong> Thisbe irenea (Lepidoptera:<br />

Riodinidae) and their effects upon attending ants. Zool. J. Linn. Soc. 94:<br />

379–393.<br />

DE VRIES, P.J. 1990. Enhancement <strong>of</strong> symbiosis between butterfly caterpillars<br />

and ants by vibrational communication. Science 248: 1104–1106.<br />

DE VRIES, P.J. 1991a. Evolutionary and ecological patterns in myrmecophilous<br />

riodinid butterflies. In: Huxley, C.R. and Cutler, D.F. (Eds) Ant-Plant<br />

Interactions, Oxford University Press, Oxford, pp. 143–156.<br />

DE VRIES, P.J. 1991b. Mutualism between Thisbe irenea butterflies and ants,<br />

and the role <strong>of</strong> ant ecology in the evolution <strong>of</strong> larval-ant associations. Biol.<br />

J. Linn. Soc. 43: 179–195.<br />

DE VRIES, P.J. and BAKER, I. 1989. Butterfly exploitation <strong>of</strong> a plant-ant<br />

mutualism: adding insult to herbivory. J. New York ent. Soc. 97: 332–340.<br />

DIAS, M.M. 1980. Morfologia da pupa de alguns Riodinidae brasileiros<br />

(Lepidoptera). Rev. bras. Entomol. 24: 181–191.<br />

DOWNEY, J.C. and ALLYN, A.C. 1979. Morphology and biology <strong>of</strong> the<br />

immature stages <strong>of</strong> Leptotes cassius theopus (Lucas) (Lepidoptera:<br />

<strong>Lycaenidae</strong>). Bull. Allyn Mus. 55: 1–27.<br />

DOWNEY, J.C. and ALLYN, A.C. 1980. Eggs <strong>of</strong> Riodinidae. J. Lepid. Soc.<br />

34: 133–145.<br />

DOWNEY, J.C. and ALLYN, A.C. 1981. Chorionic sculpturing in eggs <strong>of</strong><br />

<strong>Lycaenidae</strong>. Part I. Bull. Allyn Mus. 61: 1–29.<br />

DOWNEY, J.C. and ALLYN, A.C. 1984. Chorionic sculpturing in eggs <strong>of</strong><br />

<strong>Lycaenidae</strong>. Part II. Bull. Allyn Mus. 84: 1–44.<br />

DURDEN, C.J. and ROSE, H. 1978. <strong>Butterflies</strong> from the middle Eocene: the<br />

earliest occurrence <strong>of</strong> fossil Papilionoidea (Lepidoptera). Pearce-Sellards<br />

Series, Texas Memorial Museum 29: 1–25.<br />

HARVEY, D.J. 1987. The higher classification <strong>of</strong> the Riodinidae (Lepidoptera).<br />

Ph.D. dissertation, University <strong>of</strong> Texas, Austin, 216pp.<br />

HARVEY, D.J. 1989. Perforated cupola organs on larvae <strong>of</strong> Euselasiinae<br />

(Riodinidae). J. Lepid. Soc. 34: 127–132.<br />

HARVEY, D.J.andCLENCH,H.K.1980. Dianesia, a new genus <strong>of</strong> Riodinidae<br />

from the West Indies. J. Lepid. Soc. 34: 127–132.<br />

HARVEY, D.J. and GILBERT, L.E. 1988. Ant association, larval and pupal<br />

morphology <strong>of</strong> the neotropical Riodinid butterfly, Pandemos palaeste.<br />

Unpublished manuscript.<br />

HEWITSON, W.L. 1852–1872. Illustrations <strong>of</strong> New Species <strong>of</strong> Exotic<br />

<strong>Butterflies</strong>. Five volumes, van Voorst, London, v + 124, iv + 124, iv + 124,<br />

iii + 118, iv + 127pp., 60, 60, 60, 60 and 60 plates.<br />

HEWITSON, W.L. 1863–1878. Illustrations <strong>of</strong> Diurnal Lepidoptera:<br />

<strong>Lycaenidae</strong>, in eight parts. Van Voorst, London, x + 229 pp., 108 plates.<br />

HORVITZ, C, TURNBULL, C. and HARVEY, D.J. 1987 The biology <strong>of</strong><br />

immature Eurybia elvina (Lepidoptera: Riodinidae), a myrmecophilous<br />

metalmark butterfly. Ann. Ent. Soc. Amer. 80: 513–519.<br />

HUTCHINGS, R.W. 1991. Dinâmica de tres comunidades de Papilionoidea<br />

(Insecta: Lepidoptera) em fragmentos de floresta na Amazônia central.<br />

M.Sc. dissertation, Instituto Nacional de Pesquisas da Amazônia/Fundac,ao<br />

Universidade de Amazonas, xii + 65pp.<br />

JANZEN, D. 1984. No park is an island: increase in interference from outside<br />

as park size increases. Oikos 41: 402–410.<br />

JANZEN, D. 1986. The eternal external threat. In: Soulé, M.E. (Ed.),<br />

<strong>Conservation</strong> <strong>Biology</strong>: the Science <strong>of</strong> Scarcity and Diversity, Sinauer,<br />

Sunderland, Mass., pp. 286–303.<br />

JOHNSON, K. 1989. Revision <strong>of</strong> Chlorostrymon Clench and description <strong>of</strong><br />

two new austral Neotropical species (<strong>Lycaenidae</strong>). J. Lepid. Soc. 43:<br />

120–146.<br />

JOHNSON, K. 1991a. Neotropical Hairstreak <strong>Butterflies</strong>: Genera <strong>of</strong> the<br />

'CalycopislCalystryma grade' <strong>of</strong> Eumaeini (Lepidoptera, <strong>Lycaenidae</strong>,<br />

Theclinae) and their diagnostics. Repts Mus. Nat. Hist., Univ. Wisc.,<br />

Stevens Point 21: iv + 128pp.<br />

JOHNSON, K. 1991b. Cladistics and biogeography <strong>of</strong> two trans-Caribbean<br />

hairstreak butterfly genera, Nesiotrymon and Tetra (Lepidoptera,<br />

<strong>Lycaenidae</strong>). Amer. Mus. Novitates 3011: 1–43.<br />

JOHNSON, K. 1992. Genera and species <strong>of</strong> the Neotropical 'Elfin'-like


hairstreak butterflies (Lepidoptera, <strong>Lycaenidae</strong>, Theclinae). Repts Mus.<br />

Nat. Hist. Univ. Wisc. (Stevens Point) 22: 1–279 (2 vols).<br />

JOHNSON, S. 1985. Culturing a detritivore, Calycopis isobeon (Butler &<br />

Druce). News Lepid. Soc. 41–42.<br />

KENDALL, R.O. 1976. Larval foodplants and life history notes for some<br />

metalmarks (Lepidoptera: Riodinidae) from Mexico and Texas. Bull.<br />

Allyn Mus. 32: 1–12.<br />

LEWIS, H.L. 1973. <strong>Butterflies</strong> <strong>of</strong> the World. Follett, Chicago, xviii + 312pp.<br />

LOVEJOY, T.E., BIERREGAARD, R.O., RYLANDS, A.B., MALCOLM,<br />

J.R., QUINTELA, C.E., HARPER, L.H., BROWN JR., K.S., POWELL,<br />

A.H., POWELL, G.V.N., SCHUBART, H.O.R. and HAYS, M.B. 1986<br />

Edge and other effects <strong>of</strong> isolation on Amazon forest fragments. In: Soulé,<br />

M.E. (Ed.) <strong>Conservation</strong> <strong>Biology</strong>: the Science <strong>of</strong> Scarcity and Diversity,<br />

Sinauer, Sunderland, Mass., pp. 257–285.<br />

MALICKY, H. 1970. New aspects on the association between Lycaenid<br />

larvae (<strong>Lycaenidae</strong>) and ants (Formicidae, Hymenoptera). J. Lepid. Soc.<br />

24: 190–202.<br />

DE LA MAZA, R.R. 1988. Mariposas Mexicanas. Fondo de Cultura<br />

Economica, Mexico, 302pp.<br />

MCALPINE, W.S. 1971. A revision <strong>of</strong> the butterfly genus Calephelis<br />

(Riodinidae). J. Res. Lep. 10: 1–125.<br />

MILLER, L.D. 1970. Multiple capture <strong>of</strong> Caria ino melicerta (Riodinidae) at<br />

light. J. Lepid. Soc. 24: 13–15.<br />

MILLER, L.D. 1980. A review <strong>of</strong> the Erora laeta-group, with description <strong>of</strong><br />

a new species (<strong>Lycaenidae</strong>). J. Lepid. Soc. 34: 209–216.<br />

MILLER, L.D. and MILLER, J.Y. 1972. A new riodinid from northern<br />

Argentina (Riodinidae). Bull. Allyn Mus. 4: 1–5.<br />

MONTEIRO, R.F. 1991. Cryptic larval polychromatism in Rekoa marius<br />

Lucas and R. palegon Cramer (<strong>Lycaenidae</strong>: Theclinae). J. Res. Lepid. 29:<br />

77–85.<br />

NABOKOV, V. 1945. Notes on Neotropical Plebejinae (<strong>Lycaenidae</strong>:<br />

Lepidoptera). Psyche 52: 1–61, 8 plates.<br />

NICOLAY, S.S. 1971a. A review <strong>of</strong> the genus Arcas with descriptions <strong>of</strong> new<br />

species (<strong>Lycaenidae</strong>, Strymonini). J. Lepid. Soc. 25: 87–108.<br />

NICOLAY, S.S. 1971b. A new genus <strong>of</strong> hairstreak from Central and South<br />

America (<strong>Lycaenidae</strong>, Theclinae).J. Lepid. Soc. 25 (Suppl. 1): 1–39.<br />

NICOLAY, S.S. 1976. A review <strong>of</strong> the Hubnerian genera Panthiades and<br />

Cycnus (<strong>Lycaenidae</strong>: Eumaeini). Bull. Allyn Mus. 35: 1–30.<br />

NICOLAY, S.S. 1977. Studies on the genera <strong>of</strong> American hairstreaks. 4. A<br />

new genus <strong>of</strong> hairstreak from Central and South America (<strong>Lycaenidae</strong>:<br />

Eumaenini). Bull. Allyn Mus. 44: 1–24.<br />

NICOLAY, S.S. 1980. The genus Chlorostrymon and a new subspecies <strong>of</strong> C.<br />

simaethis. J. Lepid. Soc. 34: 253–256.<br />

NICOLAY, S.S. 1982. Studies in the genera <strong>of</strong> American hairstreaks. 6. A<br />

review <strong>of</strong> the Hubnerian genus Olynthus (<strong>Lycaenidae</strong>: Eumaeini). Bull.<br />

61<br />

Allyn Mus. 74: 1–30.<br />

ROBBINS, R.K. 1980. The Lycaenid 'false head' hypothesis: historical<br />

review and quantitative analysis.J. Lepid. Soc. 34: 194–208.<br />

ROBBINS, R.K. 1982. How many butterfly species? News Lepid. Soc. 40–41.<br />

ROBBINS, R.K. 1985. Independent evolution <strong>of</strong> 'false head' behaviour in<br />

Riodinidae. J. Lepid. Soc. 39: 221–225.<br />

ROBBINS, R.K. 1987. Evolution and identification <strong>of</strong> the New World<br />

hairstreak butterflies (<strong>Lycaenidae</strong>: Eumaeini): Eliot's Trichonis section<br />

and Trichonis Hewitson. J. Lepid. Soc. 49: 138–157.<br />

ROBBINS, R.K. 1988. Comparative morphology <strong>of</strong> the butterfly foreleg coxa<br />

and trochanter (Lepidoptera) and its systematic implications. Proc. Ent.<br />

Soc. Wash. 90: 137–154.<br />

ROBBINS, R.K. 1991. Evolution, comparative morphology, and identification<br />

<strong>of</strong> the Eumaeine butterfly genus Rekoa Kaye (<strong>Lycaenidae</strong>: Theclinae).<br />

Smith. Contr. Zool. 498, 64pp.<br />

ROBBINS, R.K. 1993. (in preparation). (Manuscript on larval foodplants <strong>of</strong><br />

Neotropical Eumaeini).<br />

ROBBINS, R.K. and AIELLO, A. 1982. Foodplant and oviposition records<br />

for Panamanian <strong>Lycaenidae</strong> and Riodinidae. J. Lepid. Soc. 36: 65–75.<br />

ROBBINS, R.K. and SMALL, G.B. 1981. Wind dispersal <strong>of</strong> Panamanian<br />

hairstreak butterflies (Lepidoptera: <strong>Lycaenidae</strong>) and its evolutionary<br />

significance. Biotropica 13: 308–315.<br />

ROBBINS, R.K. and VENABLES, B.A.B. 1991. Synopsis <strong>of</strong> a new Neotropical<br />

hairstreak genus, Janthecla, and description <strong>of</strong> new species (<strong>Lycaenidae</strong>).<br />

J. Lepid. Soc. 45: 11–33.<br />

ROSS, G.N. 1964. Life history studies on Mexican butterflies. III. Early<br />

stages <strong>of</strong> Anatole rossi, a new myrmecophilous metalmark (Lepidoptera:<br />

Riodinidae). J. Res. Lep. 3: 87–94.<br />

ROSS, G.N. 1966. Life history studies on Mexican butterflies. IV. The<br />

ecology and ethology <strong>of</strong> Anatole rossi, a myrmecophilous metalmark<br />

(Lepidoptera: Riodinidae). Ann. Ent. Soc. Amer. 59: 985–1004.<br />

ROTHSCHILD, M., NASH, R.J. and BELL, E.A. 1986. Cycasin in the<br />

endangered butterfly Eumaeus atala florida. Phytochemistry 25:<br />

1853–1854.<br />

SCHREMMER, F. 1978. Zur Bionomie und Morphologie der myrmecophilen<br />

Raupe und Puppae von der neotropischen Tagfalter-Art Hamearis erostratus<br />

(Lepidoptera: Riodinidae). Entomologica Germanica 4: 113–121.<br />

SEITZ, A. 1916–1920. Die Grossschmetterlinge der Erde 5: Die<br />

Amerikanischen Tagfalter. pp. 617–738 ('Erycinidae'); pp. 744–812<br />

('<strong>Lycaenidae</strong>', written by M. Draudt), Plates 121–159, 110A, 113B, 193,<br />

Stuttgart.<br />

SMART, P.M. 1975. The International Butterfly Book. Thomas Crowell,<br />

N.Y., 274pp.<br />

TALBOT, G. 1928. List <strong>of</strong> Rhopalocera collected by Mr. C.L. Collenette in<br />

Matto Grosso, Brazil. Bull. Hill Mus. 2: 192–220.


Threatened <strong>Lycaenidae</strong> <strong>of</strong> South Africa<br />

Michael J. SAMWAYS<br />

Department <strong>of</strong> Zoology and Entomology, University <strong>of</strong> Natal, Pietermaritzburg 3200, South Africa<br />

Southern African geology and geography<br />

To appreciate the conservation biology <strong>of</strong> lycaenids in South<br />

Africa, it is necessary to introduce the characteristic geology<br />

and geography <strong>of</strong> the area. About 180 million years B.P., the<br />

great supercontinent <strong>of</strong> Pangaea began to split. By 135 million<br />

years B.P., the southern tip <strong>of</strong> Africa looked very much in<br />

outline as it does today. Since that time, uplifting and erosion<br />

has led to the appearance <strong>of</strong> 22 physiographic regions, each<br />

characterised by altitude and surface form (Figure 1). These<br />

regions fall naturally into two groups. The first group, <strong>of</strong> 12<br />

regions, is the Interior Plateau. The second group, <strong>of</strong> 10 regions,<br />

is the Marginal Zone, which is divided from the Interior Plateau<br />

by the great divide known as the Great Escarpment.<br />

Figure 1. The 22 physiographic regions <strong>of</strong> South Africa.<br />

1. Upper Karoo; 2. Highveld; 3. Kaap Plateau; 4. Southern Kalahari;<br />

S. Bushmanland; 6. Namaqualand Highlands; 7. Bushveld Basin; 8. Bankveld;<br />

9. Pietersburg Plateau; 10. Waterberg Plateau; 11. Soutpansberg; 12. Lesotho<br />

Tableland; 13. Lebombo Hills; 14. Lowveld; 15. Middelveld; 16. Eastern<br />

Midlands; 17. Winterberg Mountains; 18. Great Karoo; 19. Doring Karoo;<br />

20. Cape Folded Mountains; 21. Little Karoo; 22. Coastal Belt.<br />

Sub regions: NN = Northern Natal; SN = Southern Natal; TR = Transkei;<br />

EC = Eastern Cape; SC = Southern Cape; WC = Western Cape.<br />

62<br />

The Interior Plateau is the southern tip <strong>of</strong> the great African<br />

Plateau, and its altitude varies from a minimum <strong>of</strong> slightly<br />

under 900m in the Kalahari Desert, to almost 3500m in the<br />

Lesotho Highlands. Here, the thick Karoo sediments <strong>of</strong> the<br />

Carboniferous to Triassic ages were covered in the Jurassic by<br />

thick lava flows, which, on cooling to a hard layer <strong>of</strong> basalt,<br />

protected the underlying s<strong>of</strong>ter rocks from weathering, giving<br />

rise to the high mountains <strong>of</strong> southern Africa.<br />

The Marginal Zone between the Great Escarpment and the<br />

coast, varies in width from 60km in the west to 240km in the east.<br />

Its elevation varies from sea-level to a maximum <strong>of</strong> 2300m in the<br />

Swartberg Range south <strong>of</strong> the Great Escarpment. In turn, the<br />

Great Escarpment is composed <strong>of</strong> several distinct mountain<br />

ranges, giving southern Africa a distinctive and varied topography.<br />

Off the coast, the northward flowing cold Benguela Current<br />

moves up the west coast, and the southward flowing warm<br />

Agulhas Current flows down the east coast. These flow patterns,<br />

along with topography and global wind patterns, influence the<br />

area's rainfall regimes. There are three distinct rainfall regions<br />

in southern Africa: summer, winter and all-season rainfall<br />

areas. The varied topography and climate has resulted in nine<br />

climatic zones (Figure 2).<br />

Species richness and threatened species<br />

The long period <strong>of</strong> geographical isolation in southern Africa,<br />

lack <strong>of</strong> glaciation, varied topography and climate has produced<br />

a wide range <strong>of</strong> biotypes and high levels <strong>of</strong> endemism and<br />

species richness in both plants and animals (Huntley 1989).<br />

Vári and Kroon (1986) list 8300 species <strong>of</strong> Lepidoptera in<br />

southern Africa (i.e. south <strong>of</strong> the Zambezi River), while Pinhey<br />

(1975) estimates that the final total will exceed 10,000.<br />

In South Africa (i.e. south <strong>of</strong> the Limpopo River), 632<br />

species <strong>of</strong> butterfly (Papilionoidea and Hesperioidea) so far<br />

have been described. Of these, 102 (16%) are under some level<br />

<strong>of</strong> threat (Henning and Henning 1989). If subspecies are included,<br />

the total number <strong>of</strong> threatened taxa is 14%.<br />

Vari and Kroon (1986) list 389 lycaenid species in southern<br />

Africa. Of these, 310 occur in South Africa, and 105 species and


Fig. 2. The nine climatic zones in South Africa.<br />

1. Subtropical Lowveld; 2. Subtropical Coast; 3. Temperate Coast;<br />

4. Mediterranean; 5. Plateau Slopes; 6. Temperate Eastern Plateau;<br />

7. Subtropical Plateau; 8. Semi-arid Plateau; 9. Desert.<br />

subspecies are in some way threatened (Henning and Henning<br />

1989) (Table 1 and Figure 3). This is 75% <strong>of</strong> all threatened<br />

butterflies in the country. Two species <strong>of</strong> lycaenid are now<br />

Extinct, and another one species and one subspecies are<br />

Endangered (Table 1 and Figure 3).<br />

Particularly significant is that 96% (101) <strong>of</strong> the threatened<br />

species (71% species and 25% subspecies) are endemic to<br />

South Africa (Clark and Dickson 1971; Henning and Henning<br />

1989; Murray 1935). These figures are high for a portion <strong>of</strong> a<br />

continental land mass, but not unusual for other biotic groups in<br />

the subcontinent. Of the remaining four species, only one is<br />

widespread in Africa, while the other three have limited<br />

distributions or are on the edge <strong>of</strong> a range just extending into<br />

South Africa.<br />

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

species<br />

Figure 4 illustrates that by far the richest area for rare endemics<br />

is the Cape Fold Mountains (physiographic areas 19,20 and 21<br />

in Figure 1). Most occur on mountain slopes, but some inhabit<br />

ridges (e.g. Poecilmitis wykehami, Thestor dicksoni dicksoni),<br />

peaks (e.g. Lepidochrysops outeniqua, P. endymion, P. balli)<br />

Figure 3. Status <strong>of</strong> South Africa's 105 threatened lycaenid species and subspecies using the <strong>of</strong>ficial <strong>IUCN</strong> categories (Data from Henning and Henning, 1989).<br />

63


Table 1. Status <strong>of</strong> threatened lycaenid species and subspecies in South Africa (from the South African Red Data Book, Henning and Hcnning 1989).<br />

Lipteninae<br />

Alaena margaritacea Eltringham<br />

Deloneura immaculata Trimen<br />

Durbania amakosa albescens Quickelberge<br />

Durbania amakosa flavida Quickelberge<br />

Ornipholidotos peucetia penningtoni (Riley)<br />

Liphyrinae<br />

Aslauga australis Cottrell<br />

Miletinae<br />

Thestor brachycerus (Trimen)<br />

Theslor compassbergae Quickelberge & McMaster<br />

Theslor dicksoni calviniae Riley<br />

Thestor dicksoni dicksoni Riley<br />

Thestor dryburghi Van Son<br />

Thestor kaplani Dickson & Stephen<br />

Thestor montanus pictus Van Son<br />

Thestor pringlei Dickson<br />

Thestor rossouwi Dickson<br />

Thestor stepheni Swanepoel<br />

Thestor strutti Van Son<br />

Thestor swanepoeli Pennington<br />

Thestor tempe Pennington<br />

Thestor yildizae Koçak<br />

Theclinae<br />

Aloeides caledoni Tite & Dickson<br />

Aloeides carolynnae Dickson<br />

Aloeides clarki Tite & Dickson<br />

Aloeides dentatis dentatis (Swierstra)<br />

Aloeides dentatis maseruna (Riley)<br />

Aloeides egerides (Riley)<br />

Aloeides kaplani Tite & Dickson<br />

Aloeides lutescens Tite & Dickson<br />

Aloeides merces Henning & Henning<br />

Aloeides nollothi Tite & Dickson<br />

Aloeides nubilus Henning & Henning<br />

Aloeides pringlei Tite & Dickson<br />

Aloeides rossouwi Henning & Henning<br />

Aloeides trimeni southeyae Tite & Dickson<br />

Argyrocupha malagrida cedrusmontana<br />

Dickson & Stephen<br />

Argyrocupha malagrida malagrida (Wallengren)<br />

Argyrocupha malagrida maryae Dickson & Henning<br />

Argyrocupha malagrida paarlensis (Dickson)<br />

Bowkeria phosphor borealis Quickelberge<br />

Bowkeria phosphor phosphor (Trimen)<br />

Capys penningtoni Riley<br />

Chrysoritis oreas (Trimen)<br />

Chrysoritis cottrelli Dickson<br />

Erikssonia acraeina Trimen<br />

Hypolycaena lochmophila Tite<br />

Iolaus (Epamera) aphnaeoides Trimen<br />

Iolaus (Epamera) diametra natalica Vári<br />

Iolaus (Pseudiolaus) lulua Riley<br />

Oxychaeta dicksoni (Gabriel)<br />

Phasis pringlei Dickson<br />

V<br />

Ex<br />

R<br />

I<br />

R<br />

R<br />

I<br />

I<br />

R<br />

R<br />

I<br />

R<br />

I<br />

R<br />

I<br />

I<br />

R<br />

I<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

I<br />

R<br />

I<br />

R<br />

R<br />

I<br />

R<br />

I<br />

R<br />

R<br />

R<br />

V<br />

V<br />

R<br />

R<br />

R<br />

R<br />

R<br />

E<br />

V<br />

I<br />

R<br />

R<br />

R<br />

V<br />

R<br />

64<br />

Phasis thero cedarbergae Dickson & Wykeham<br />

Poecilmitis adonis Pennington<br />

Poecilmitis aureus Van Son<br />

Poecilmitis azurius Swanepoel<br />

Poecilmitis balli Dickson & Henning<br />

Poecilmitis brooksi learei Dickson<br />

Poecilmitis daphne Dickson<br />

Poecilmitis endymion Pennington<br />

Poecilmitis henningi Bampton<br />

Poecilmitis hyperion Dickson<br />

Poecilmitis irene Pennington<br />

Poecilmitis kaplani Henning<br />

Poecilmitis lyncurium (Trimen)<br />

Poecilmitis lyndseyae Henning<br />

Poecilmitis nigricans nigricans (Aurivillius)<br />

Poecilmitis nigricans zwartbergae Dickson<br />

Poecilmitis orientalis Swanepoel<br />

Poecilmitis pan Pennington<br />

Poecilmitis penningtoni Riley<br />

Poecilmitis pyramus Pennington<br />

Poecilmitis pyroeis hersaleki Dickson<br />

Poecilmitis rileyi Dickson<br />

Poecilmitis stepheni Dickson<br />

Poecilmitis swanepoeli Dickson<br />

Poecilmitis trimeni Riley<br />

Poecilmitis wykehami Dickson<br />

Trimenia wallengrenii (Trimen)<br />

Polyommatinae<br />

Anthene minima (Trimen)<br />

Cyclyrius babaulti (Stempffer)<br />

Lepidochrysops bacchus Riley<br />

Lepidochrysops badhami Van Son<br />

Lepidochrysops balli Dickson<br />

Lepidochrysops hypopolia (Trimen)<br />

Lepidochrysops jamesi classensi Dickson<br />

Lepidochrysops jamesi jamesi Swanepoel<br />

Lepidochrysops jefferyi (Swierstra)<br />

Lepidochrysops littoralis Swanepoel & Vari<br />

Lepidochrysops loewensteini (Swanepoel)<br />

Lepidochrysops lotana Swanepoel<br />

Lepidochrysops methymna dicksoni Tite<br />

Lepidochrysops oosthuizeni Swanepoel & Vari<br />

Lepidochrysops oreas oreas Tite<br />

Lepidochrysops outeniqua Swanepoel & Vari<br />

Lepidochrysops penningtoni Dickson<br />

Lepidochrysops pephredo (Trimen)<br />

Lepidochrysops poseidon Pringle<br />

Lepidochrysops pringlei Dickson<br />

Lepidochrysops quickelbergei Swanepoel<br />

Lepidochrysops swanepoeli Pennington<br />

Lepidochrysops titei Dickson<br />

Lepidochrysops victori Pringle<br />

Lepidochrysops wykehami Tite<br />

Orachrysops ariadne (Butler)<br />

Orachrysops niobe (Trimen)<br />

Tuxentius melaena griqua (Trimen)<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

I<br />

R<br />

R<br />

R<br />

I<br />

R<br />

R<br />

R<br />

Ex<br />

R<br />

R<br />

R<br />

R<br />

R<br />

V<br />

E<br />

R<br />

R<br />

R<br />

R<br />

R<br />

I<br />

R<br />

R<br />

R<br />

I<br />

R<br />

I<br />

R<br />

V<br />

R


Figure 4. Distribution <strong>of</strong> South Africa's threatened lycaenid taxa across the 22 physiographic regions. (N.B. A few taxa occur in more than one region).<br />

and gullies (e.g. P. azurius). Some species, despite the harsh<br />

winter conditions <strong>of</strong> the Cape, can be found at high elevations<br />

(e.g. Argyrocupha malagrida cedrusmontana at 1900m and<br />

Aloeides pringlei at 2072m ). The lower elevations <strong>of</strong> the Cape<br />

(physiographic subregions 22SC and 22WC) are also rich in<br />

localised endemics (Figure 4) illustrating the wide range <strong>of</strong><br />

biotopes occupied by this group <strong>of</strong> butterflies in the<br />

topographically varied and geologically stable area on the<br />

southern tip <strong>of</strong> Africa.<br />

The next richest area in rare, threatened lycaenids is the east<br />

coast and hinterland. Most <strong>of</strong> the blue butterflies in these areas<br />

are fairly abundant, and the mountain peaks do not have the<br />

same richness <strong>of</strong> localised endemics as in the Cape. One species<br />

(Lepidochrysops lowensteini) occurs at about 2800m in the<br />

Lesotho Tableland, while the middle altitudes support 14<br />

threatened taxa. The lower altitudes <strong>of</strong> the east coast are also<br />

relatively rich in threatened endemics. When the four east coast<br />

subregions <strong>of</strong> physiographic region 22 are added together<br />

(Figure 4), the coastal plain from Mozambique in the north to<br />

Port Elizabeth in the south hosts 13 taxa.<br />

The Central Plateau (regions 1, 2, 3, 4, and 5) and the<br />

Northern Transvaal savannah (regions 7, 8, 9, 10, 11, 13, and<br />

14) are poor in localised, threatened endemics, with regions 4,<br />

5, and 13 supporting none. The reason for this is uncertain, but<br />

with the monotonous landscape and little opportunity for<br />

topographic isolation, it is possible that competition from more<br />

vigorous species has taken place. This may even have happened<br />

in recent times with Lepidochrysops hypopolia. One <strong>of</strong> the two<br />

original localities for this species is Potchefstroom on the<br />

Highveld, where it was found in 1879: today the habitat is<br />

occupied by the closely-related L. praeterita, with L. hypopolia<br />

not having been recorded for over 110 years.<br />

65<br />

Threats leading to taxon extinction<br />

Overcollecting<br />

There appear to be no verified cases <strong>of</strong> butterfly species going<br />

extinct through overcollecting (New 1984, Pyle et al. 1981).<br />

Many <strong>of</strong> the lycaenids in South Africa are in remote rugged<br />

localities which are not readily accessible. Some <strong>of</strong> the lowland<br />

species are more easily collected, but there is no verified case<br />

<strong>of</strong> overcollecting affecting a population level permanently.<br />

Invasive species<br />

Many plant species <strong>of</strong> the Cape fynbos heaths and shrublands<br />

have their seeds dispersed and buried by ants. At the turn <strong>of</strong> the<br />

century, the exotic Argentine ant (Iridomyrmex humilis (Mayr))<br />

appeared in South Africa and recently has invaded the fynbos<br />

and has begun to supplant indigenous ants (Bond and Slingsby<br />

1984). For blue butterflies <strong>of</strong> the Cape this is a serious<br />

development for two reasons. Firstly, I. humilis does not<br />

disperse and bury the seeds as do the native ants: this will<br />

eventually lead to the alteration <strong>of</strong> the habitat through loss <strong>of</strong><br />

Cape Proteaceae species by gradual attrition <strong>of</strong> seed reserves.<br />

Secondly, because many lycaenids depend directly on certain<br />

indigenous ant species as hosts, the loss <strong>of</strong> native ants must<br />

inevitably have serious long-term repercussions for the survival<br />

<strong>of</strong> many <strong>of</strong> the lycaenids.<br />

By late 1978, I. humilis had reached the Highveld, but to<br />

date has not been recorded from Natal. Should further increase<br />

in the range <strong>of</strong> I. humilis occur, which is probable, many<br />

lycaenids throughout the country could be affected. This may


e inevitable as there are no known methods <strong>of</strong> arresting the<br />

expansion orcontrolling the population levels <strong>of</strong> this aggressive,<br />

invasive ant.<br />

Apart from the very serious threat from the Argentine ant,<br />

none <strong>of</strong> the other ant species or other hosts associated with<br />

lycaenids are under threat. In fact, most species are widespread<br />

and abundant (e.g. Anoplolepis custodiens (Smith),<br />

Crematogaster spp., Camponotus spp.).<br />

The invasive spread <strong>of</strong> alien trees and shrubs has also been<br />

significant for most South African biomes (Macdonald et al.<br />

1986). For blue butterflies, genera such as Pinus, Acacia,<br />

Eucalyptus, Solarium and Rubus pose the greatest threat by<br />

partial or total alteration <strong>of</strong> the habitat. Without containment,<br />

these weeds are likely to be increasingly threatening to lycaenid<br />

populations (Figure 5). The habitat <strong>of</strong> Argyrocupha malagrida<br />

malagrida is already being heavily invaded by alien vegetation,<br />

as is the habitat <strong>of</strong> one <strong>of</strong> the populations <strong>of</strong> Poecilmitis pan.<br />

There is no evidence that classical biological control<br />

programmes have had any adverse effect upon lycaenid<br />

conservation, or even upon any other threatened insect species<br />

in South Africa (Samways 1988).<br />

Change in landscape use<br />

Loss <strong>of</strong> habitat is a complex issue and involves more than<br />

simply loss <strong>of</strong> natural areas to agriculture and urbanisation. For<br />

example, the increased subdivision <strong>of</strong> the landscape by vehicle<br />

tracks and roads prevents the spread <strong>of</strong> natural fires which are<br />

an important natural influence preventing the succession from<br />

grassland to scrub (Tainton and Mentis 1984). The savannah is<br />

burnt to simulate the effects <strong>of</strong> lightning strikes and maintain<br />

habitats in a relatively pristine state. Conversely, accidental<br />

fires that are more intense or regular than would normally be the<br />

case in nature can be a threat in their own right. This is<br />

particularly the case adjacent to urban or commercially forested<br />

areas. Such a fire threat, as well as direct habitat loss, faces, for<br />

example, Thestor yildizae on Table Mountain in Cape Town.<br />

The majority <strong>of</strong> taxa are included in the butterfly Red Data<br />

Book (Henning and Henning 1989) because <strong>of</strong> their extremely<br />

limited natural distribution (Figure 5) particularly in<br />

mountainous areas. However, agricultural and urban<br />

developments are major threats for many lycaenids, particularly<br />

those at low elevations. In the case <strong>of</strong> Orachrysops ariadne,<br />

Figure 5. Major threats, in decreasing magnitude, to South Africa's 105 rare lycaenid taxa. (N.B. Some lycaenids are threatened by more than one<br />

category.)<br />

Many are not threatened but simply have strictly limited distributions. All are threatened by global warming and increased radiation from the southern hole in<br />

the ozone layer. 'Extinct' is the <strong>IUCN</strong> categorisation; other categories are the actual threats. It is not known what the threats were to the Extinct species. (Data<br />

calculated from identified and listed threats for all red-listed species in Henning and Henning, 1989).<br />

66


although its site is privately protected, absence <strong>of</strong> large herbivores<br />

and lax management <strong>of</strong> vegetation is posing a threat to the longterm<br />

security <strong>of</strong> the species.<br />

Many <strong>of</strong> the threat categories are interrelated, e.g. urban<br />

expansion and fire hazards or dam construction. The overall<br />

root cause is the same: increased human population pressure<br />

which is particularly threatening to the lowland species but not<br />

so much to those <strong>of</strong> the rugged, relatively inaccessible, Cape<br />

mountains.<br />

Global warming and thinning <strong>of</strong> the ozone layer<br />

In the long term, these threats may outweigh all others. The<br />

earth's temperature may rise by 3°C by the year 2050, and<br />

models predict that although there will be little change at the<br />

equator, the poles may be TC warmer (Pearman 1988). Besides<br />

a rise in sea level, there may also be a shift in seasons. For South<br />

Africa, projections are for a generally warmer, drier situation,<br />

where soil moisture conditions in the Highveld might be 11–18%<br />

drier than at present (Huntley et al. 1989). The exceptionally<br />

species-rich Cape fynbos would be particularly affected, with<br />

the possible risk <strong>of</strong> a collapsing domino effect with many<br />

localised invertebrates disappearing along with their plant<br />

hosts. Additive upon these general effects will be an increase in<br />

inclement conditions such as droughts, hailstorms and<br />

hurricanes. Adding further stress will be the growing hole in the<br />

ozone layer <strong>of</strong> the stratosphere over the South Pole, with<br />

depletion in parts <strong>of</strong> up to 50% (Brunke 1988). South Africa,<br />

with its southerly geographical position, is likely to be<br />

particularly affected by the increased levels <strong>of</strong> ultraviolet<br />

radiation. Already, in 1989, there was a 45% depletion<br />

(Scourfield et al. 1990).<br />

All these adverse environmental pressures bode ill for many<br />

lycaenids with their localised distributions and specialised lifecycles<br />

(many <strong>of</strong> which are still to be determined). Temperature<br />

falls by 0.6° C for every 100m rise in altitude, which suggests<br />

that to maintain the same local thermal environment, species<br />

must move up the mountainside by at least 500m by the middle<br />

<strong>of</strong> next century. Such movement would be difficult for species<br />

whose habitat has been partitioned by roads, buildings, etc.<br />

(Siegfried 1989). Further, as many <strong>of</strong> the South African lycaenids<br />

occur on isolated ridges and peaks, they cannot go higher and<br />

must adapt or die. In the short term, genetic adaptation is<br />

unlikely, and these ecological pressures may be too great for<br />

survival.<br />

<strong>Conservation</strong> measures<br />

The production <strong>of</strong> the Red Data Book on butterflies (Henning<br />

and Henning 1989) has been an important step forward<br />

(Samways 1989a). Although not all the species and subspecies<br />

listed may be truly threatened (apart from the effects <strong>of</strong> significant<br />

global events) with many further populations awaiting discovery,<br />

the book has focused on the rarity <strong>of</strong> many butterflies, particularly<br />

67<br />

lycaenids, and is a major stepping stone for further research.<br />

Few insects are protected by law in South Africa, and each<br />

province has its own ordinances. To date no specific species or<br />

subspecies is protected by law in Natal or the Orange Free State.<br />

In the Cape Province, according to Ordinance 19 <strong>of</strong> 1974,<br />

Schedule 2, Protected Wild Animals may not be hunted, killed,<br />

captured or kept in captivity without a permit from the<br />

Department <strong>of</strong> Nature and Environmental <strong>Conservation</strong>.<br />

Amendment <strong>of</strong> Schedule 2 (13 February 1976) gives full<br />

protection to: Aloeides egerides, A. lutescens, Argyrocupha<br />

malagrida malagrida, Trimenia wallengrenii, Oxychaeta<br />

dicksoni, Lepidochrysops bacchus, Poecilmitis endymion, P.<br />

lyncurium, P. nigricans nigricans, P. rileyi, Thestor dicksoni<br />

dicksoni and T. kaplani. In the Transvaal, by Ordinance 12 <strong>of</strong><br />

1983, Section 45 (Schedule 7) (Protected Wild Animals),<br />

Poecilmitis aureus is protected.<br />

The 582 nature reserves in southern Africa total 7 x 10 6 ha,<br />

5.8% <strong>of</strong> the land surface (Huntley 1989). Siegfried (1989)<br />

estimated that 74% <strong>of</strong> vascular plant species and over 90% <strong>of</strong><br />

each <strong>of</strong> the vertebrate groups are represented in nature reserves.<br />

Such comparative data is not available for insects, but <strong>of</strong> the<br />

threatened lycaenids listed by Henning and Henning (1989),<br />

two are extinct, and only 24 (23%) occur in nature reserves or<br />

wilderness areas, all but one state owned. Apart from the<br />

protection ordinances, this means that 75% are not<br />

geographically protected in nature reserves. For many <strong>of</strong> these,<br />

apart from global atmospheric threats, they are relatively safe as<br />

their localities are in remote terrain. Given the extent <strong>of</strong> remote<br />

terrain there is also the possibility <strong>of</strong> the occasional new species<br />

being found from time to time.<br />

In recent years there has been an increase in invertebrate<br />

conservation awareness in South Africa. Aloeides dentatis<br />

dentatis in particular has become quite a celebrity as the<br />

Roodepoort City Council has established the 12ha Ruimsig<br />

Entomological Reserve specifically for the butterfly (Henning<br />

and Henning 1985).<br />

Although such authorities as the National Parks Board, the<br />

Natal Parks Board, the Defence Force, the Universities <strong>of</strong><br />

Natal, Pretoria and Stellenbosch have a strong interest in<br />

conservation and support invertebrate conservation projects,<br />

only the Transvaal Provincial Administration has a full-time<br />

Invertebrate <strong>Conservation</strong> Officer. This is encouraging, but<br />

still inadequate bearing in mind that there are about 80,000<br />

described species <strong>of</strong> insect in South Africa (Prinsloo 1989) and<br />

that this may only be a quarter <strong>of</strong> the total percentage, including<br />

a large proportion <strong>of</strong> endemics. The emphasis to date has been<br />

on the conservation <strong>of</strong> large vertebrates, while the Wildlife<br />

Society principally supports the conservation <strong>of</strong> habitats and<br />

whole natural areas. It is well known that vertebrates are not<br />

good indicator or umbrella species for invertebrate conservation.<br />

Other invertebrates with specific habitat requirements and<br />

appropriately sized home ranges are better as flag species<br />

(Samways 1989b).<br />

Although the setting aside <strong>of</strong> game and nature reserves has<br />

given refuge to the great majority <strong>of</strong> vertebrates, it applies to<br />

less than one-quarter <strong>of</strong> the blue butterflies. Nature reserves


nevertheless are playing a role. Acquisition <strong>of</strong> much more land<br />

for state-owned nature reserves is unlikely, as most land is<br />

firmly accounted for. This means that invertebrate zoologists,<br />

entomologists and naturalists must monitor areas where blue<br />

butterfly populations are still maintaining a foothold, and lobby<br />

for conservation <strong>of</strong> that patch <strong>of</strong> land. As these habitat patches<br />

are indeed small (tennis court size in the case <strong>of</strong> Argyrocupha<br />

malagrida malagrida) it is a feasible proposition, as shown by<br />

the Henning brothers and the Roodepoort City Council in the<br />

setting aside <strong>of</strong> the Ruimsig Entomological Reserve.<br />

Summary<br />

Southern Africa, with its long stable geological history and<br />

highly varied topography and climate, is rich in insect species.<br />

In South Africa, 632 species <strong>of</strong> butterfly have been described,<br />

<strong>of</strong> which 102 (16%) are under some sort <strong>of</strong> threat. There are 389<br />

lycaenid species in southern Africa, with 310 species occurring<br />

in South Africa. A total <strong>of</strong> 105 species and subspecies <strong>of</strong><br />

lycaenid are threatened, 75% <strong>of</strong> all threatened butterfly taxa in<br />

the country. Of the 105 threatened species two lycaenid species<br />

are Extinct; one species and one subspecies are Endangered;<br />

seven taxa are Vulnerable; 71 are Rare; and 23 are Indeterminate.<br />

A high proportion <strong>of</strong> the threatened lycaenid taxa (71% species<br />

+ 25% subspecies = 96% total) are endemic. Most <strong>of</strong> the<br />

threatened taxa occur in the Cape fold mountain area, usually on<br />

mountain slopes, and sometimes at high elevations (up to<br />

1,800m). Many <strong>of</strong> the rare lycaenids also occur along the east<br />

coast and hinterland, including one at 2800m. The Central<br />

Plateau and Northern Transvaal savannah are poor in localised<br />

endemics.<br />

Overcollecting is not a threat. Of great seriousness is the<br />

invasive Argentine ant (Iridomyrmex humilis) which is<br />

supplanting native ant hosts for lycaenids, the key species in<br />

maintaining the character <strong>of</strong> the habitats through seed burial.<br />

Invasive exotic plant species are also a threat for several<br />

lycaenids, as is fire, mining, dam building, forestry, and, for one<br />

species, neglectful management. For the lowland species,<br />

agricultural and urban expansion are the greatest threats, but for<br />

many <strong>of</strong> the species, which generally inhabit mountainous<br />

areas which are rugged or remote, it is simply that they have an<br />

extremely limited distribution.<br />

Of great concern in southern Africa is the overall effect <strong>of</strong><br />

global warming and the hole in the ozone layer. Projections<br />

indicate that most species will need to shift in elevation by<br />

500m by the year 2050 if they are to remain at the same<br />

temperature. For most species, especially those that inhabit<br />

hilltops, such a shift would be ecologically and genetically<br />

impossible.<br />

Twelve lycaenid taxa are protected by law in the Cape<br />

Province, and one species in the Transvaal. None are protected<br />

in the Orange Free State or Natal. Only 23% <strong>of</strong> threatened<br />

lycaenids occur within nature reserves or wilderness areas,<br />

compared with over 90% for each <strong>of</strong> the vertebrate groups and<br />

68<br />

74% for vascular plants. One species, Aloeides dentatis, has<br />

been allocated a 12ha reserve <strong>of</strong> its own. Further acquisition <strong>of</strong><br />

patches <strong>of</strong> land for specific lycaenid populations is a feasible<br />

short-term approach to further the cause <strong>of</strong> invertebrate<br />

conservation.<br />

Acknowledgements<br />

The Foundation for Research Development and the University<br />

<strong>of</strong> Natal Research Fund provided financial assistance. Messrs.<br />

Stephen and Graham Henning kindly criticised the text and Mrs<br />

Ann Best and Mrs Myriam Preston kindly processed the<br />

manuscript.<br />

References<br />

BOND, W. and SLINGSBY, P. 1984. Collapse <strong>of</strong> an ant-plant mutualism: The<br />

Argentine ant (Iridomyrmex humilis) and myrmecochorous Proteaceae.<br />

Ecology 65: 1031–1037.<br />

BRUNKE, G.C. 1988. Tropospheric background measurements <strong>of</strong> CFCI3(F-<br />

II) conducted at Cape Point, South Africa, since 1979. In: Macdonald,<br />

I.A.W. and Crawford, R.J.M. (Eds), Long-term Data Series relating to<br />

southern Africa's Renewable Natural Resources. South African National<br />

Scientific Programmes Report No. 157. Foundation for Research<br />

Development, Council for Scientific and Industrial Research, Pretoria. pp.<br />

434–435.<br />

CLARK, G.W. and DICKSON, C.G.C. 1971. Life Histories <strong>of</strong> South African<br />

Lycaenid <strong>Butterflies</strong>. Purnell, Cape Town. 272pp.<br />

HENNING, S.F. and HENNING, G.A. 1985. South Africa's endangered<br />

butterflies. Quagga 10: 16–17.<br />

HENNING, S.F. and HENNING, G.A. 1989. South African Red Data Book-<br />

<strong>Butterflies</strong>. South African National Scientific Programmes Report No.<br />

158. Foundation for Research Development, Council for Scientific and<br />

Industrial Research, Pretoria. 175pp.<br />

HUNTLEY, B.J. (Ed.) 1989. Biotic Diversity in Southern Africa: Concepts<br />

and <strong>Conservation</strong>. Oxford University Press, Cape Town. 380pp.<br />

HUNTLEY, B.J., SIEGFRIED, R. and SUNTER, C. 1989. South African<br />

Environments into the 21st Century. Human and Rousseau Tafelberg,<br />

Cape Town. 127pp.<br />

MACDONALD, I.A.W., KRUGER, F.J. and FERRAR, A.A. (Eds) 1986. The<br />

Ecology and Management <strong>of</strong> Biological Invasions in southern Africa.<br />

Oxford University Press, Oxford. 324pp.<br />

MURRAY, D.P. 1985. South African <strong>Butterflies</strong>. A Monograph <strong>of</strong> the Family<br />

<strong>Lycaenidae</strong>. Staples Press, London. 195pp.<br />

NEW, T.R. 1984. Insect <strong>Conservation</strong> – An Australian Perspective. Junk,<br />

Dordrecht. 184pp.<br />

PEARMAN, G.I. (Ed.) 1988. Greenhouse: Planning for Climate Change.<br />

CSIRO, Melbourne.<br />

PINHEY, E.C.G. 1975. Moths <strong>of</strong> Southern Africa. Tafelberg, Cape Town.<br />

273pp.<br />

PRINSLOO, G.L. 1989. Insect identification services in South Africa.<br />

Proceedings <strong>of</strong> the Seventh Entomological Congress organized by the<br />

Entomological Society <strong>of</strong> southern Africa, Pietermaritzburg 10–13 July<br />

1989. p.107.<br />

PYLE, R.M., BENTZIEN, M.M. and OPLER, P.A. 1981. Insect conservation.<br />

Ann. Rev. Entomol. 26: 233–258.<br />

SAMWAYS, M.J. 1988. Classical biological control and insect conservation:<br />

Are they compatible? Env. Conserv. 15: 349–354.<br />

SAMWAYS, M.J. 1989a. Amnesty for insects. S. Afr. J. Sci. 85: 571–572.


SAMWAYS, M.J. 1989b. Insect conservation and landscape ecology: A casehistory<br />

<strong>of</strong> bush crickets (Tettigoniidae) in southern France. Env. Cons. 16:<br />

217–226.<br />

SCOURFIELD, M.W.J., BODEKER, G., BARKER, M.D., DIAB, R.D. and<br />

SALTER, L.F. 1990. Ozone: the South African connection. S. Afr. J. Sci.<br />

86: 279–281.<br />

SIEGFRIED, W.R. 1989. Preservation <strong>of</strong> species in southern African nature<br />

reserves. In: Huntley, B.J. (Ed.) Biotic Diversity in Southern Africa:<br />

69<br />

Concepts and <strong>Conservation</strong>.<br />

TAINTON, N.M. and MENTIS, M.T. 1984. Fire in grassland. In: Booysen, P.<br />

de V. and Tainton, N.M. (Eds). Ecological Effects <strong>of</strong> Fire in South African<br />

Ecosystems. Springer-Verlag, Berlin. pp. 115–147.<br />

VARI, L. and KROON, D. 1986. Southern African Lepidoptera: A Series <strong>of</strong><br />

Cross-referenced Indices. Lepidopterists' Society <strong>of</strong> southern Africa and<br />

the Transvaal Museum, Pretoria. 198pp.


Introduction<br />

Australian <strong>Lycaenidae</strong>: conservation concerns<br />

T.R. NEW<br />

Department <strong>of</strong> Zoology, La Trobe University, Bundoora, Victoria 3083, Australia<br />

The <strong>Lycaenidae</strong> <strong>of</strong> Australia comprise about 140 described<br />

species (Table 1, Figure 1). The family is most diverse in the<br />

northern tropical parts <strong>of</strong> the country, where the fauna <strong>of</strong><br />

northern Queensland has strong relationships with that <strong>of</strong> New<br />

Guinea and parts <strong>of</strong> the Oriental Region. Indeed, Eliot (1973)<br />

chose to follow Gressitt (1956) in considering the northern tip<br />

<strong>of</strong> Queensland, Cape York Peninsula, as part <strong>of</strong> the Oriental<br />

Region in delimiting areas for considering lycaenid distribution<br />

patterns.<br />

Taxonomic appraisal <strong>of</strong> the fauna, at least <strong>of</strong> the adult<br />

stages, is relatively complete (Common and Waterhouse 1981)<br />

and it is possible to provide a reasonably sound appraisal <strong>of</strong> the<br />

status and broad scale distribution <strong>of</strong> all species. Dunn and<br />

Dunn (1991) give maps <strong>of</strong> the documented distribution <strong>of</strong><br />

Australian butterflies. A few further species undoubtedly remain<br />

to be discovered, but most future changes in taxonomy are<br />

likely to be made at the species/subspecies interface, with some<br />

subspecies being elevated in status as more information becomes<br />

available. Not surprisingly for a large continent extending from<br />

the humid tropics to cold temperate regions and with large arid<br />

and semiarid zones, the distribution <strong>of</strong> many lycaenid species is<br />

circumscribed, and there is room for considerably more research<br />

to clarify the precise ranges and status <strong>of</strong> many <strong>of</strong> the more<br />

elusive resident taxa.<br />

Only one riodinine (Praetaxila segecia punctaria<br />

(Fruhstorfer)) extends into Australia, where it is confined to the<br />

Table 1. Taxonomic summary <strong>of</strong> the Australian <strong>Lycaenidae</strong> (data from<br />

Dunn and Dunn 1991).<br />

Subfamily<br />

Liphyrinae<br />

Riodininae<br />

Theclinae<br />

Polyommatinae<br />

Total<br />

No. genera<br />

1<br />

1<br />

15<br />

22<br />

39<br />

No. species<br />

1<br />

1<br />

75<br />

65<br />

142<br />

70<br />

extreme north <strong>of</strong> Queensland; its close relatives are found in<br />

New Guinea. Intriguingly, four species <strong>of</strong> Lycaeninae occur in<br />

New Zealand, but none in Australia. In general, the Australian<br />

<strong>Lycaenidae</strong> show some attenuation in diversity from those <strong>of</strong><br />

New Guinea and Malaysia, and the number <strong>of</strong> subfamilies is<br />

smaller. This is particularly obvious at the tribal level: Australia<br />

has 32 species <strong>of</strong> Luciini compared with about a hundred in<br />

New Guinea, and only four Arhopalini compared with 38 in<br />

New Guinea and slightly more than a hundred in Malaysia. Of<br />

the generally widespread subfamilies four are absent: Miletinae;<br />

Curetinae; Lycaeninae; and Poritiinae. The two predominant<br />

subfamilies in Australia are Polyommatinae andTheclinae, and<br />

endemism is high in both (Polyommatinae: 4/22 genera, 22/65<br />

species; Theclinae: 6/15 genera, 25/75 species). Theclini, in<br />

particular, are a major radiation in the Australian butterfly<br />

fauna (Kitching 1981). These figures for endemism do not<br />

include numerous putative subspecies which are clearly restricted<br />

to Australia and <strong>of</strong>ten have very limited distributions. The<br />

biological status <strong>of</strong> most <strong>of</strong> these is by no means clear and some,<br />

at least, may constitute sibling species complexes.<br />

Many endemic taxa are very rare. Several, indeed, are<br />

known only from one or two localities. For example,<br />

Hypochrysopspiceatus Kerr, McQueen & Sands has only been<br />

found in two small colonies in southern Queensland and only<br />

one <strong>of</strong> these now remains. Jalmenus aridus Graham & Moulds,<br />

one <strong>of</strong> very few butterflies believed to be endemic to Australia's<br />

interior arid zone, is known from one colony in Western<br />

Australia. A number <strong>of</strong> subspecies <strong>of</strong> Ogyris Westwood and<br />

Candalides Hiibner, inter al., also have very limited distributions.<br />

Some <strong>of</strong> these, such as the 'O. idmo group' in Western Australia,<br />

are taxonomically more complex than currently documented<br />

(Field 1992), and perhaps <strong>of</strong> greater conservation concern than<br />

suspected at present.<br />

It is notable that ant-dependence to differing extents occurs<br />

in virtually all the genera which are diverse and widespread.<br />

This habit may have been instrumental in leading to<br />

diversification <strong>of</strong> <strong>Lycaenidae</strong> in some semiarid regions where<br />

plant growth is very irregular and the spectrum <strong>of</strong> food plants<br />

limited. Larval feeding habits differ considerably between<br />

various genera. Additional foodplant records continue to be<br />

accumulated (e.g. Valentine and Johnson 1988). In Ogyris


Figure 1. Australia: main political regions and lycaenid fauna.<br />

Regions denoted by inilial letters: ACT, Australian Capital Territory; NSW, New South Wales; NT, Northern Territory; SA, South Australia; T, Tasmania;<br />

V, Victoria; WA, Western Australia. Lycaenid subfamilies: L, Liphyrinae; P, Polyommatinae; R, Riodininae; T, Theclinae. Figures are no. <strong>of</strong> included genera/<br />

species.<br />

Figure 2. Australia: main specific places mentioned in the text.<br />

71


caterpillars <strong>of</strong> some species may live in ant nests throughout<br />

their lives.<br />

By contrast many other Ogyris larvae feed on mistletoes<br />

(Loranthaceae, Viscaceae), and this unusual host plant group<br />

may have been a major basis for speciation in this genus. In<br />

contrast, Hypochrysops has exploited a taxonomically diverse<br />

range <strong>of</strong> flora, so that diversification has occurred in these<br />

genera by using contrasting ecological strategies. However,<br />

very few lycaenids feed directly on one dominant tree genus,<br />

Eucalyptus.<br />

The Australian fauna<br />

Some highlights <strong>of</strong> the Australian fauna are described below<br />

with much <strong>of</strong> the information gleaned from the extensive data<br />

summarised by Common and Waterhouse (1981).<br />

• The widespread oriental species Liphyra brassolis major<br />

Rothschild is rather rare in northern tropical Australia,<br />

where it occurs in association with Oecophylla ants.<br />

• In the Luciini, many taxa are very local. For example, Lucia<br />

limbaria Swainson occurs in isolated localities in the south<br />

and east <strong>of</strong> mainland Australia.<br />

• Several species <strong>of</strong> Acrodipsas Sands (formerly referred to<br />

Pseudodipsas C. & R. Felder) are also rare and localised: A.<br />

arcana (Miller & Edwards) is known from only one hilltop<br />

in New South Wales, and A. hirtipes Sands from a hilltop<br />

near Coen (northern Queensland) (Figure 2) where it occurs<br />

with A. melania Sands. The latter is known also from a<br />

single specimen captured at the very tip <strong>of</strong> Cape York<br />

Peninsula. A. brisbanensis cyrilus (Anderson & Spry) is a<br />

local subspecies in Victoria and A. b. brisbanensis (Miskin),<br />

although more widely distributed along the east coast, is<br />

also regarded as rare. Indeed, none <strong>of</strong> the seven species <strong>of</strong><br />

Acrodipsas is common.<br />

• Paralucia Waterhouse & Turner contains three species <strong>of</strong><br />

which two are rare: P. spinifera Edwards & Common (see<br />

Dexter and Kitching, this volume) is one <strong>of</strong> our rarest<br />

lycaenids, and is known only from one restricted area <strong>of</strong><br />

New South Wales; and P. pyrodiscus lucida Crosby has<br />

recently aroused considerable conservation interest in<br />

Victoria (see New 1992, and New, this volume).<br />

• Hypochrysops C. & R. Felder is the most speciose lycaenid<br />

genus in Australia, and many <strong>of</strong> the 18 species are rare and<br />

local (Sands 1986 and Sands, this volume). It is absent from<br />

Tasmania, and most <strong>of</strong> the species are northern or east<br />

central in distribution. Several New Guinea (or closely<br />

related) species are known only from northern Cape York,<br />

and the only species in Western Australia is H. halyaetus<br />

Hewitson. As with Philiris Rober and Arhopala Boisduval,<br />

Australian subspecies <strong>of</strong> some New Guinea species have<br />

developed.<br />

• Ogyris, with 12 endemic Australian species, also occurs in<br />

New Guinea but is <strong>of</strong> considerable biological interest in<br />

Australia. Several species include a number <strong>of</strong> named<br />

72<br />

infraspecific taxa and many <strong>of</strong> these are local and rare forms<br />

attractive to collectors. Both subspecies <strong>of</strong> O. idmo Hewitson<br />

are extremely rare, as are O. otanes C. & R. Felder, O.<br />

ianthus Waterhouse, and O. iphis Waterhouse & Lyell. All<br />

merit strenuous conservation measures, in common with<br />

several more widespread forms.<br />

• Jalmenus Hubner contains 10 species. J. pseudictinus is<br />

regarded as very local in parts <strong>of</strong> eastern Queensland, as is<br />

J. lithochroa Waterhouse in southern South Australia. J.<br />

dementi Druce is known from a very small region <strong>of</strong><br />

northwestern Australia, and J. aridus from a colony near<br />

Kalgoorlie (Figure 2).<br />

• The single species <strong>of</strong> Pseudalmenus Druce, P. chlorinda<br />

(Blanchard) is confined to southeastern Australia, and seven<br />

subspecies have been described. Four <strong>of</strong> these are from<br />

Tasmania (see Prince, this volume), where populations<br />

separated by only a few kilometres have very different wing<br />

markings. P. c. fisheri Tindale, from Western Victoria, is<br />

also very local.<br />

• Some other lycaenids are not regarded as rare but are very<br />

clearly restricted to particular geographical regions or<br />

ecological communities: Neolucia hobartensis (Miskin),<br />

with two subspecies, is an alpine/subalpine species <strong>of</strong><br />

southeastern Australia.<br />

Distributional and diversity patterns<br />

The distribution patterns (Dunn and Dunn 1991) reveal several<br />

trends relevant to the consideration <strong>of</strong> lycaenid conservation in<br />

Australia. The northern part <strong>of</strong> Queensland, as for many other<br />

biota, supports a large number <strong>of</strong> taxa on the southernmost<br />

fringes <strong>of</strong> their Oriental/New Guinea distributions. Remnant<br />

rainforests in this region are particularly important butterfly<br />

habitats (Monteith and Hancock 1977) and the individual<br />

patches may be viewed as an 'archipelago' <strong>of</strong> these in northern<br />

Queensland.<br />

Lycaenid diversity is greatest around the eastern and<br />

southeastern fringe <strong>of</strong> the Australian mainland, with a smaller<br />

number <strong>of</strong> species in the west or southwest. Very few species<br />

occur in the climatically inhospitable inland.<br />

A very high proportion <strong>of</strong> Australian <strong>Lycaenidae</strong> are forest<br />

and open woodland-frequenting taxa. A few herb/shrub<br />

community taxa, such as Lampides boeticus (L.), Theclinesthes<br />

serpentata (Herrich-Schaffer) and Zizina labradus (Godart),<br />

are amongst the most widely distributed lycaenids in Australia.<br />

Habitat relationships in the southeast are exemplified by an<br />

analysis for the Australian Capital Territory (Table 2: Kitching<br />

et al. 1978).<br />

Most <strong>of</strong> the non-endemic species belong to oriental genera,<br />

and essentially constitute the 'younger northern element' <strong>of</strong> the<br />

fauna. Endemism at the generic level is distinctively more<br />

southern (and, especially, southeastern), with the implication<br />

that some, at least, may represent speciation from earlier<br />

colonisers than those taxa which occur solely in the north; a


Table 2. Habitat relationships <strong>of</strong> 25 species <strong>of</strong> I.ycaenidac recorded from<br />

the Australian Capital Territory. Figures given are numbers <strong>of</strong> species.<br />

Data from Kitching et al. 1978.<br />

No. Habitat<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Lowland savannah<br />

Savannah woodland<br />

Dry sclerophyll forest<br />

Wet sclerophyll forest<br />

Alpine zone<br />

Total<br />

5<br />

16<br />

15<br />

11<br />

4<br />

No. species<br />

shared with habitat<br />

2 3 4 5<br />

5 5<br />

12<br />

2<br />

5<br />

6<br />

1<br />

1<br />

2<br />

4<br />

Not<br />

shared<br />

similar situation occurs in Australian Satyrinae (New in press).<br />

Couchman and Couchman (1977) considered Tasmanian forms<br />

<strong>of</strong> P. chlorinda to be 'evidently <strong>of</strong> very ancient origin'. Centres<br />

<strong>of</strong> endemism, local 'critical faunas', can thus be delimited with<br />

some degree <strong>of</strong> reliability.<br />

<strong>Conservation</strong><br />

Lycaenid diversity is highest in those parts <strong>of</strong> Australia which<br />

are subject to substantial, rapid and largely irreversible changes<br />

by European people. Thus diversity is highest in the eastern and<br />

southeastern fringe <strong>of</strong> the Australian mainland, with smaller<br />

numbers <strong>of</strong> species in the west or southwest and very few in the<br />

climatically inhospitable inland. Environmental changes through<br />

human activities have accelerated in recent decades and show<br />

little sign <strong>of</strong> abating or slowing in the near future: assessing<br />

threats to taxa at both local and national levels becomes essential<br />

if the loss <strong>of</strong> species and notable subspecies is to be avoided.<br />

For some parts <strong>of</strong> the country it is not possible to determine<br />

if some taxa were formerly more widespread than they are at<br />

present. Inferences on past endangering processes, including<br />

habitat change and loss, thus contain a large element <strong>of</strong> historical<br />

supposition.<br />

As stated earlier, local 'critical faunas' can be delimited<br />

with some degree <strong>of</strong> reliability. In addition, there are a number<br />

<strong>of</strong> 'critical habitats' for <strong>Lycaenidae</strong> which can be identified.<br />

These occur on a macroscale (isolated pockets <strong>of</strong> rainforest – as<br />

in northern Queensland, alpine or mallee vegetation) and also<br />

as specific sites which support one or more narrow endemics<br />

and which may be especially vulnerable–several <strong>of</strong> the hilltops<br />

referred to in the taxonomic summary are good examples <strong>of</strong><br />

this. Because <strong>of</strong> their restricted or sporadic distributions over<br />

this vast island continent, a substantial number <strong>of</strong> endemic<br />

lycaenids on the Australian mainland must be considered<br />

vulnerable to continuing habitat change, even though they are<br />

not directly or imminently threatened.<br />

–<br />

4<br />

2<br />

3<br />

–<br />

73<br />

Threatened taxa<br />

In a preliminary survey <strong>of</strong> threatened insects in Australia, Hill<br />

and Michaelis(1988) included nine species <strong>of</strong> <strong>Lycaenidae</strong>, five<br />

<strong>of</strong> which were known only from one or two sites. The survey<br />

was based on a questionnaire circulated widely to entomologists<br />

in Australia, seeking information on priorities for insect<br />

conservation. In all, 24 taxa <strong>of</strong> <strong>Lycaenidae</strong> (Table 3) were noted<br />

by respondents and these included three subspecies <strong>of</strong> P.<br />

chlorinda, two <strong>of</strong> which occurred in Tasmania. Tasmanian taxa<br />

were noted as <strong>of</strong> concern by other respondents, so that<br />

conservation concern for native <strong>Lycaenidae</strong> has a broad<br />

geographic base within Australia.<br />

It is notable that the Hill and Michaelis list includes threatened<br />

<strong>Lycaenidae</strong> from all mainland states except the Northern<br />

Territory, where documentation is relatively incomplete.<br />

Nadolny (1987) noted that the most endangered butterfly in<br />

New South Wales is likely to be Paralucia spinifera, which is<br />

Table 3. Taxa <strong>of</strong> <strong>Lycaenidae</strong> in Australia which may be threatened.<br />

Data from Hill and Michaelis (1988), digested from respondents to survey<br />

conducted by Australian National Parks and Wildlife Service, closing February<br />

1985. States abbreviated as in Figure 1.<br />

Taxon<br />

Acrodipsas arcana (Miller & Edwards)<br />

A. brisbanensis brisbanensis (Miskin)<br />

A. illidgei (Waterhouse & Lyell)<br />

A. n.sp.<br />

A. myrmecophila (Waterhouse & Lyell)<br />

Hypochrysops apollo Miskin<br />

H. clean Grose-Smith<br />

H. epicurus Miskin<br />

H. hippuris Hewitson<br />

H. ignitus ignitus (Leach)<br />

H. piceatus Kerr, Macqueen & Sands<br />

Jalmenus lithochroa Waterhouse<br />

Jamides cytus claudia (Waterhouse & Lyell)<br />

Ogyris amaryllis meridionalis Bethune-Baker<br />

O. idmo halmaturia Tepper<br />

O. idmo idmo Hewitson<br />

O. otanes C. & R. Felder<br />

Paralucia spinifera Edwards & Common<br />

Philiris azula Wind & Clench<br />

P. ziska titeus D'Abrera<br />

Pseudalmenus chlorinda chlorinda (Blanchard)<br />

P. c. barringlonensis Waterhouse<br />

P. c. conara Couchman<br />

Theclinesthes albocincta (Waterhouse)<br />

State(s)<br />

Q,NSW<br />

Q.NSW<br />

Q<br />

WA<br />

Q,NSW,V<br />

Q<br />

Q<br />

Q,NSW<br />

Q<br />

SA,NSW,V<br />

Q<br />

SA<br />

0<br />

SA<br />

SA,V<br />

WA<br />

SA,NSW,WA<br />

NSW<br />

0<br />

Q<br />

T<br />

NSW<br />

T<br />

SA


<strong>of</strong> considerable phylogenetic importance in possibly linking<br />

Paralucia with related genera. P. spinifera, the Bathurst Copper,<br />

has recently been the subject <strong>of</strong> more detailed study in the State<br />

(Dexter and Kitching, this volume).<br />

However, in general there are few detailed accounts <strong>of</strong><br />

range contractions <strong>of</strong> taxa other than on a very local scale or by<br />

inferences from older collectors who claim that some species<br />

are now scarcer or more restricted than in the past. More survey<br />

work, particularly in poorly documented areas, is needed to<br />

identify all threatened taxa <strong>of</strong> the Australian <strong>Lycaenidae</strong>.<br />

Major threats to Australian lycaenids<br />

The survey by Hill and Michaelis (1988) identified some <strong>of</strong> the<br />

major threats to Australian lycaenids and these are listed below,<br />

together with other threats specified in the literature:<br />

• land clearing, sometimes by fire, for agriculture and urban<br />

development;<br />

• urbanisation and tourist resort development;<br />

• pest fly control;<br />

• roadworks;<br />

• mining;<br />

• collecting (individual and commercial);<br />

• agricultural practices, e.g. pasture improvement.<br />

Examples <strong>of</strong> species threatened by the ecological processes<br />

follow; others can be found in the literature. Any <strong>of</strong> these<br />

species might also be susceptible to overcollecting.<br />

(i) The best known and most accessible site for H. piceatus<br />

is a small patch <strong>of</strong> mistletoe-bearing Casuarina along some<br />

200m <strong>of</strong> road verge, which could be eliminated easily and<br />

inadvertently by road widening.<br />

(ii) O. otanes, recently (July 1989) the first butterfly to be<br />

nominated for listing under the Victorian Flora and Fauna<br />

Guarantee Act, is known in Victoria from a single hilltopping<br />

site in the arid northwest <strong>of</strong> the state which has been the subject<br />

<strong>of</strong> physical disturbance during the establishment <strong>of</strong> a<br />

trigonometric survey point, by vehicles on sand-dunes. Other<br />

populations <strong>of</strong> O. otanes were known in the 1970s but were<br />

destroyed by collecting.<br />

(iii) One <strong>of</strong> the very few sites at which Acrodipsas<br />

myrmecophila occurs, and the only one known in Victoria at<br />

present, is currently the target <strong>of</strong> mining exploration.<br />

Additionally, it is not clear whether the two species <strong>of</strong> Acrodipsas<br />

described from a hill near Coen breed on the hill or in nearby<br />

lowland vegetation and then hilltop. In both cases, much nearby<br />

land is currently subject to mining exploration and this could<br />

pose a threat. Several such hilltops in Australia are among the<br />

classic collecting localities favoured by butterfly collectors,<br />

and such rare species could be rendered additionally vulnerable<br />

by uncontrolled collecting.<br />

(iv) The several species associated with coastal mangrove<br />

vegetation are vulnerable as drainage occurs for urban or tourist<br />

resort development.<br />

(v) The detailed appraisal <strong>of</strong> P. chlorinda in Tasmania<br />

(Couchman and Couchman 1977), which has recently been<br />

updated (Prince 1988), claims that it has been eliminated over<br />

74<br />

whole areas where one or other <strong>of</strong> its required major resources<br />

(a eucalypt, an Acacia – normally A. dealbata – and an<br />

Iridomyrmex ant) have been destroyed. Couchman and<br />

Couchman located the species in more than 50 localities in the<br />

years following 1945 but at the time <strong>of</strong> writing their account<br />

found it difficult to think <strong>of</strong> more than 10 localities where P.<br />

chlorinda might then survive. 'Pasture improvement' with<br />

removal <strong>of</strong> all mature eucalypts and acacias in paddocks badly<br />

affected P. c. conara, and most <strong>of</strong> the habitat <strong>of</strong> the eastern P.<br />

c. chlorinda has been clearfelled for woodchips. Two other<br />

forms are extinct, one because <strong>of</strong> local tree clearing and one<br />

because <strong>of</strong> housing development and clearing/burning, and<br />

some existing forms are highly localised (see Prince, this<br />

volume).<br />

(vi) Of the three species <strong>of</strong> Acrodipsas listed as threatened<br />

by Hill and Michaelis (1988), A. arcana is threatened by fires<br />

and clearing, A. illidgei (a mangrove-frequenting species, see<br />

Samson this volume) by urbanisation and use <strong>of</strong> insecticides<br />

against mosquitoes, and a third (undescribed) species is subject<br />

to habitat destruction by clearing for agriculture.<br />

Geographical areas <strong>of</strong> importance for lycaenids<br />

While it is essential to identify threatened species as foci for<br />

conservation attention, some larger geographical areas also<br />

merit particular attention by harbouring diverse or unusual<br />

faunal groups. Worthy <strong>of</strong> note are:<br />

i) Cape York Peninsula. This is a major region <strong>of</strong> faunal<br />

interchange between Australia and New Guinea, with many<br />

northern <strong>Lycaenidae</strong> not extending further into Australia. This<br />

northern element includes some 75 species (more than half the<br />

Australian <strong>Lycaenidae</strong>) and, whereas a number <strong>of</strong> these extend<br />

further down the east coast or elsewhere, many species are both<br />

rare and restricted to the 'far north', <strong>of</strong>ten to forested areas. The<br />

number <strong>of</strong> rainforest <strong>Lycaenidae</strong> decreases latitudinally from<br />

about 32 species at Iron Range to none in the cool temperate<br />

forests <strong>of</strong> Victoria and Tasmania. Tropical rainforests are a key<br />

habitat for maintaining the integrity <strong>of</strong> tropical fauna in<br />

Australia.<br />

ii) Southern Queensland. Kitching (1981) highlights the<br />

very high diversity <strong>of</strong> <strong>Lycaenidae</strong> in southern Queensland, and<br />

suggests that this reflects the confluence <strong>of</strong> northern and southern<br />

faunas. Thus, many <strong>of</strong> the northern taxa reach their southern<br />

limits here and the putatively older southern forms, their most<br />

northern range extensions. There are also a number <strong>of</strong> rare<br />

species found only in this region, which supports around half<br />

the Australian lycaenid species.<br />

iii) Western Victoria, particularly the Grampians<br />

('Gariwerd') mountains and the 'Mallee region'. A number <strong>of</strong><br />

rare inland forms occur in these areas, and some taxa from<br />

semiarid regions are scarce or non-existent elsewhere.<br />

These regions include substantial areas <strong>of</strong> National Park or<br />

other reserves: Iron Range and the Grampians ('Gariwerd')<br />

National Parks are two examples. In situations where the prime<br />

conservation need is security <strong>of</strong> habitat and our state <strong>of</strong><br />

knowledge <strong>of</strong> particular species does not permit intervention to


markedly influence management, protection <strong>of</strong> these reserves<br />

and their enhancement is the most important step which can be<br />

taken. Presence <strong>of</strong> rare <strong>Lycaenidae</strong>, such as P. chlorinda fisheri<br />

in the Grampians ('Gariwerd') can here augment pressures for<br />

habitat protection and maintenance <strong>of</strong> the integrity <strong>of</strong> reserves.<br />

The role <strong>of</strong> insects in such planning in Australia is in its infancy,<br />

and the limited work on lycaenid conservation to date has been<br />

almost entirely species-targeted.<br />

In the current Australian political climate, demands for<br />

multiple land use, sometimes involving substantial intrusion<br />

into National Parks, are not uncommon.<br />

Legislation<br />

None <strong>of</strong> the threatened lycaenids identified by Hill and Michaelis<br />

(1988) is formally listed as endangered in any state other than<br />

Victoria and Queensland, so there is no legal mechanism for<br />

their protection other than their fortuitous and largely<br />

undocumented/unmonitored incidence in nominally safe<br />

'Reserves'.<br />

The issue <strong>of</strong> 'species listing' is a controversial one in<br />

Australia. In the past a number <strong>of</strong> insect species have been<br />

gazetted as 'protected' in various States, with little apparent<br />

reason. This practice has in some instances alienated concerned<br />

collectors and others who are in a position to help very positively<br />

with the documentation needed to clarify the status and well<br />

being <strong>of</strong> thetaxa involved. In no case until 1989 had 'listing' <strong>of</strong><br />

an insect species been accompanied by any form <strong>of</strong> formal<br />

undertaking or provision for study <strong>of</strong> the biology <strong>of</strong> the species.<br />

The single species listed in Queensland legislation,<br />

Acrodipsas illidgei (see Samson, this volume) was designated<br />

in 1990 as 'permanently protected fauna', an extraordinarily<br />

high level <strong>of</strong> nominal protection placing it on a par with some<br />

charismatic vertebrates.<br />

The recent legislation enacted in Victoria in 1988 and<br />

known as the 'Flora and Fauna Guarantee' is pioneering in<br />

scope. Taxa can be nominated for listing, and made subject to<br />

an 'interim conservation order' – a legal hiatus which provides<br />

the opportunity for a more formal appraisal <strong>of</strong> the status <strong>of</strong> the<br />

species during a period when it is nominally protected from<br />

further intensification <strong>of</strong> the threatening processes such as<br />

habitat destruction. Essentially, the onus then falls on the State<br />

Department for <strong>Conservation</strong> and Natural Resources to<br />

investigate the species, and to clarify its need for conservation.<br />

If identified as threatened, a management plan must be produced<br />

to clarify the major steps needed to protect the species. Clearly,<br />

this is limited to within Victorian State boundaries, but many<br />

conservationists are hoping that despite its Utopian and possibly<br />

impracticable (because <strong>of</strong> restricted logistic capability) ideals,<br />

the Guarantee will tangibly safeguard rare Victorian endemics<br />

and isolated remnant or outlying populations <strong>of</strong> taxa more<br />

common in other parts <strong>of</strong> the country.<br />

As noted earlier, O. otanes was the first species to be listed<br />

in Victoria and this was followed by nominations for a number<br />

<strong>of</strong> other lycaenids (Acrodipsas myrmecophila, A. brisbanensis,<br />

Paraluciapyrodiscus lucida, as further candidates. The isolated<br />

75<br />

hill on which A. myrmecophila occurs in Victoria, Mt Piper,<br />

also supports A. brisbanensis and has been listed (as Butterfly<br />

Community No. 1) for investigation as a 'threatened community',<br />

again a pioneering step for butterfly conservation in Australia,<br />

on the basis <strong>of</strong> this unique co-occurrence. Very few lycaenid<br />

species have thus been accorded consideration for legal<br />

protection in Australia.<br />

Both O. otanes and O. idmo have been placed, for some 10<br />

years, on a 'voluntary restricted collecting code' list <strong>of</strong> the<br />

Entomological Society <strong>of</strong> Victoria. This 'code', heeded by the<br />

great majority <strong>of</strong> responsible collectors, restricts the numbers<br />

<strong>of</strong> adults which can be taken by any person to two each year, and<br />

deters the collection <strong>of</strong> early stages. Because Ogyris larvae<br />

pupate in groups under the loose bark <strong>of</strong> eucalypts, this stage is<br />

undoubtedly vulnerable to overcollecting: it is the easiest one to<br />

obtain in order to procure first class cabinet specimens, and<br />

collectors tend to take a surfeit <strong>of</strong> pupae to counter losses due<br />

to parasitoids.<br />

Although particular lycaenids are highly sought after by<br />

collectors in Australia, the extent <strong>of</strong> commercial dealing in<br />

butterflies is rather low. A few local forms have appeared in<br />

dealers' lists in Australia in recent years, including various<br />

subspecies <strong>of</strong> Pseudalmenus and Ogyris. Prices have been in<br />

the order <strong>of</strong> $5–10/specimen, rarely more, and the market does<br />

not appear to be large. No information is to hand on numbers <strong>of</strong><br />

specimens sold to overseas collectors as opposed to those in<br />

Australia, but a number <strong>of</strong> the rarer taxa have been listed as 'expupa',<br />

implying their wild origin.<br />

The future for the Australian <strong>Lycaenidae</strong><br />

Habitat alteration in many parts <strong>of</strong> Australia has continued at an<br />

accelerating pace in recent years. In some documented cases<br />

habitat degradation has resulted in serious range contractions<br />

for lycaenids (Hill and Michaelis 1988; Couchman and<br />

Couchman 1977, updated by Prince 1988). There are<br />

undoubtedly other undocumented cases amongst the Australian<br />

lycaenids. There is clearly still a very long way to go to improve<br />

knowledge <strong>of</strong> biology and local distributions. Even for Victoria,<br />

perhaps the best documented mainland state, distribution maps<br />

are incomplete on a fine scale and progress towards improving<br />

this condition is slow.<br />

However, there are some encouraging signs. There is, for<br />

example, the improved legislation pioneered by Victoria, which<br />

has produced the legal mechanisms necessary for protection <strong>of</strong><br />

invertebrate species in this State and has taken this beyond mere<br />

prohibition <strong>of</strong> collecting to provision for sound scientific<br />

management. Another encouraging sign in recent years is that<br />

an increasing number <strong>of</strong> biologists and others are becoming<br />

aware <strong>of</strong> the effects <strong>of</strong> habitat alteration on insects, are seeking<br />

to counter them at all levels, and to document more fully the<br />

distribution <strong>of</strong> butterflies in Australia. Finally, there is evidence<br />

<strong>of</strong> an increasing awareness <strong>of</strong> the importance <strong>of</strong> invertebrates in<br />

ecosystems and natural community dynamics. <strong>Butterflies</strong> are


playing their part as 'invertebrate ambassadors' in promoting<br />

this awareness: the Eltham Copper (see New, this volume) has,<br />

more than any other single invertebrate species, helped to bring<br />

the plight <strong>of</strong> many insects to wide public and government<br />

attention in Victoria. A. illidgei in Queensland (see Samson,<br />

this volume) has also been a key factor in saving some important<br />

coastal areas from poorly planned development.<br />

The topic <strong>of</strong> insect conservation is now widely discussed in<br />

Australia, and many <strong>of</strong> the themes <strong>of</strong> concern are addressed by<br />

New (1984, 1992) and Greenslade and New (1991). A more<br />

general appraisal <strong>of</strong> the Australian environment and the<br />

widespread changes that have occurred during only 200 years<br />

<strong>of</strong> European settlement are included in Jeans (1986). The<br />

factors and processes exemplified above for lycaenids are <strong>of</strong><br />

much wider concern in affecting much <strong>of</strong> the endemic Australian<br />

biota.<br />

References<br />

COMMON, I.F.B. and WATERHOUSE, D.F. 1981. <strong>Butterflies</strong> <strong>of</strong> Australia.<br />

2nd ed. Angus and Robertson, Sydney.<br />

COUCHMAN, L.E. and COUCHMAN, R. 1977. The <strong>Butterflies</strong> <strong>of</strong> Tasmania.<br />

Tasmanian Year Book 1977: 66–96.<br />

DUNN, K.L. and DUNN, L.E. 1991. Review <strong>of</strong> Australian <strong>Butterflies</strong>:<br />

Distribution, Life History and Taxonomy. Part 3. <strong>Lycaenidae</strong>. Power<br />

Press, Bayswater, Victoria.<br />

ELIOT, J.N. 1973. The higher classification <strong>of</strong> the <strong>Lycaenidae</strong>: a tentative<br />

arrangement. Bull. Brit. Mus. nat. Hist. (Ent.) 28: 373–506.<br />

FIELD, R. 1992. [Research Grant Report on 'Life history studies and species<br />

determination <strong>of</strong> the Ogyris idmo Hewitson (Lepidoptera: <strong>Lycaenidae</strong>)<br />

complex in Western Australia']. Myrmecia 28 (4): 12–17.<br />

GREENSLADE, P. and NEW, T.R. 1991. Australia: conservation <strong>of</strong> a<br />

76<br />

continental insect fauna. In: Collins, N.M. and Thomas, J.A. (Eds)<br />

<strong>Conservation</strong> <strong>of</strong> Insects and their Habitats, pp. 33–70. Academic Press,<br />

London.<br />

GRESSITT, J.L. 1956. Some distribution patterns <strong>of</strong> Pacific Island Fauna.<br />

Syst. Zool. 5: 11–32.<br />

JEANS, D.N. (ed.) 1986. Australia – a Geography. Vol. 1. The Natural<br />

Environment. Sydney University Press, Sydney.<br />

HILL, L. and MICHAELIS, F.B. 1988. <strong>Conservation</strong> <strong>of</strong> insects and related<br />

wildlife. Occasional Paper No. 13. Australian National Parks and Wildlife<br />

Service, Canberra.<br />

KITCHING, R.L. 1981. The geography <strong>of</strong> the Australian Papilionoidea. In:<br />

Keast, A. (Ed.) Ecological Biogeography <strong>of</strong> Australia. W. Junk, The<br />

Hague, pp. 979–1105.<br />

KITCHING, R.L., EDWARDS, E.D., FERGUSON, D., FLETCHER, M.B.<br />

and WALKER, J.M. 1978. The butterflies <strong>of</strong> the Australian Capital<br />

Territory. J. Aust. ent. Soc. 17: 125–133.<br />

MONTEITH, G.B. and HANCOCK, D.L. 1977. Range extensions and notable<br />

records <strong>of</strong> butterflies <strong>of</strong> Cape York Peninsula, Australia. Aust. ent. Mag.<br />

4: 21–38.<br />

NADOLNY, C. 1987. Rainforest <strong>Butterflies</strong> in New South Wales: their<br />

Ecology, Distribution and <strong>Conservation</strong>. NSW National Parks and Wildlife<br />

Service, Sydney.<br />

NEW, T.R. 1984. Insect <strong>Conservation</strong>: an Australian Perspective. W. Junk,<br />

Dordrecht.<br />

NEW, T.R. 1992. <strong>Conservation</strong> <strong>of</strong> butterflies in Australia. J. Res. Lepid. 29:<br />

237–253.<br />

NEW, T.R. (in press). The evolution and characteristics <strong>of</strong> the Australian<br />

butterfly fauna. In: Jones, R.E., Kitching, R.L. and Pierce, N.E. (Eds)<br />

<strong>Biology</strong> <strong>of</strong> the Australian <strong>Butterflies</strong>. CSIRO, Melbourne.<br />

PRINCE, G.B. 1988. The conservation status <strong>of</strong> the Hairstreak Butterfly<br />

Pseudalmenus chlorinda Blanchard in Tasmania. (Report to Department<br />

<strong>of</strong> Lands, Parks and Wildlife. Hobart, Tasmania).<br />

SANDS, D.P.A. 1986. A revision <strong>of</strong> the genus Hypochrysops C. & R. Felder<br />

(Lepidoptera: <strong>Lycaenidae</strong>). Entomonograph No. 7. E.J. Brill, Leiden.<br />

VALENTINE, P.S. and JOHNSON, S.J. 1988. Some new larval food plants<br />

for north Queensland <strong>Lycaenidae</strong> (Lepidoptera). Aust. ent. Mag. 14:<br />

89–91.


PART 3. ACCOUNTS OF PARTICULAR TAXA<br />

OR COMMUNITIES<br />

The following examples complement and extend points raised<br />

in the previous section. The series <strong>of</strong> case-histories presented<br />

includes some which are classics in insect conservation and<br />

which have necessitated considerable research over many years,<br />

and some which are regarded as priorities for future study and<br />

appraisal. They range from the well-known to the speculative.<br />

Most are taxon-based, but several neotropical assemblages<br />

regarded as 'threatened communities' are included also.<br />

For two <strong>of</strong> the taxon studies which have been <strong>of</strong> critical<br />

importance in advancing the knowledge and practice <strong>of</strong> lycaenid<br />

conservation, authors were not found. Rather than ignore these<br />

and impoverish the perspective which I hope this book will<br />

provide, I have included accounts <strong>of</strong> these abstracted from<br />

published papers and reports. The 'species accounts' are,<br />

therefore at two levels: those attributed to particular authors<br />

who have usually played leading roles in the study <strong>of</strong> the<br />

particular species and non-attributed accounts which give a<br />

more historical perspective from an 'outsider'. These latter<br />

accounts may be open to update and revision.<br />

Several species are discussed in substantial detail in the<br />

previous section, and accounts <strong>of</strong> the Mission Blue and Karner<br />

Blue (see Cushman and Murphy, this volume) are <strong>of</strong> the utmost<br />

importance in insect conservation in North America. Likewise,<br />

Bálint (this volume) provides short accounts <strong>of</strong> 17 further<br />

eastern European taxa, mostly little known but which can act as<br />

foci for future attention in that region. Despite lack <strong>of</strong> current<br />

detailed information on such taxa, appraisals such as those<br />

presented for these taxa are invaluable in demonstrating the<br />

scope <strong>of</strong> regional needs. Some taxa have been especially<br />

significant in raising public awareness <strong>of</strong> butterfly conservation<br />

(New 1991). The Large Blue (Maculinea arion) in Britain and<br />

Introductory comment<br />

77<br />

the Mission Blue (Plebejus icarioides missionensis) in the<br />

United States are, perhaps, the most widely known. The Xerces<br />

Blue (Glaucopsychexerces) became extinct in California shortly<br />

before the Second World War, and is commemorated in the<br />

name <strong>of</strong> the Xerces Society, a leading body for the promotion<br />

<strong>of</strong> invertebrate conservation in North America and elsewhere.<br />

Foundation <strong>of</strong> the Society in December 1971 was, indeed,<br />

stimulated by the plight <strong>of</strong> the Large Blue in Britain (Pyle<br />

1976).<br />

Many <strong>of</strong> the most informative and influential cases <strong>of</strong><br />

lycaenid conservation in the northern hemisphere have involved<br />

conservation <strong>of</strong> subspecies, sometimes <strong>of</strong> remnant populations<br />

or those close to the edge <strong>of</strong> a species' range.<br />

As far as possible, a standard sequence <strong>of</strong> subheadings is<br />

adopted in this section, to facilitate comparison between taxa.<br />

Comments on' Status', unless otherwise made clear, refer to the<br />

geographical range or country indicated, and not necessarily<br />

the entire species range. ('Red List') denotes that the species is<br />

listed by <strong>IUCN</strong> (1990). The abbreviation 'UTM', used in<br />

several European species accounts, refers to 'Universal<br />

Transverse Mercator' projection.<br />

References<br />

<strong>IUCN</strong> 1990. <strong>IUCN</strong> Red List <strong>of</strong> Threatened Animals. <strong>IUCN</strong>, Cambridge.<br />

NEW, T. R. 1991. Butterfly <strong>Conservation</strong>. Oxford University Press, Melbourne.<br />

PYLE, R.M. 1976. <strong>Conservation</strong> <strong>of</strong> Lepidoptera in the United States. Biol.<br />

Conserv. 9: 55–75.


The mariposa del Puerto del Lobo<br />

Agriades zullichi Hemming (= nevadensis Zullich)<br />

M.L. MUNGUIRA and J. MARTIN<br />

Departamento de Biologia (Zoologia), Facultad des Ciencìas, Universidad Autonoma de Madrid, Madrid, Spain<br />

Country: Spain (southeast).<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

This species is probably one <strong>of</strong> Europe's rarest butterflies.<br />

Restricted to the high altitude schist screes <strong>of</strong> the Sierra Nevada<br />

(Granada Province), it is threatened in part <strong>of</strong> its range by a<br />

planned redevelopment <strong>of</strong> the ski station already built within its<br />

range. Its collection has only been reported five times although<br />

the area in which it lives is a classical collecting site. Nevertheless,<br />

a recent survey (Munguira 1989) found it was still abundant in<br />

one locality.<br />

Taxonomy and Description: The species was regarded as a<br />

subspecies <strong>of</strong> Agriades glandon (de Prunner) for almost 50<br />

years after it was described as a new species. This fact has<br />

prevented it from being listed separately by Heath (1981) and<br />

Viedma and Gomez (1976, 1985).<br />

Recent studies consider it to be a true species (Kudrna 1986;<br />

Munguira 1989) based on features <strong>of</strong> larval and adult<br />

78<br />

morphology, geographical isolation and its distinct ecology.<br />

There are two other species <strong>of</strong> this genus in Europe including A.<br />

glandon which is common in the Pyrenees, Alps and<br />

Scandinavian tundra and A. pyrenaicus Boisduval which lives<br />

in several high altitude ranges in southern Europe. These two<br />

species have been listed as 'endemic' (Viedma and Gomez<br />

1976, 1985) and 'vulnerable' (Heath 1981) respectively.<br />

Distribution: A. zullichi has only been found in three localities<br />

in the Sierra Nevada (southern Spain), each one in a different 10<br />

x 10km UTM square (Munguira 1989). One suitable area for the<br />

species is a 40km long and 5–10km wide area on the higher<br />

altitudes <strong>of</strong> the Sierra, but the butterfly is only present in small<br />

scattered patches due to the distribution <strong>of</strong> its foodplant. A<br />

thorough study <strong>of</strong> the whole area has not been made, and all the<br />

available information comes from two classic sites (Puerto del<br />

Lobo and Veleta). The altitudinal range <strong>of</strong> the species is 2600<br />

to 2900m.<br />

Population Size: Population numbers seem to be high in the<br />

type locality (Puerto del Lobo) where at least several hundred<br />

adults were collected in 1968 (Fernandez-Rubio 1970). In a<br />

larval survey from this locality we recorded 56 larvae in 0.5ha<br />

giving a rough estimate <strong>of</strong> 3000 butterflies for the total population<br />

in the area. On the other hand, the most endangered colony in<br />

the Veleta probably only supports c. 100 adults, and in a thorough<br />

larval survey we only found 12 larvae.<br />

Habitat and Ecology: The species is restricted to schist screes<br />

on wind-exposed hill ridges where the vegetation cover is poor.<br />

The climatic vegetation is a grassland <strong>of</strong> the Erigeronto frigidi<br />

– Festuceto clementei series. The foodplant (Vitaliana<br />

primuliflora) grows in tight cushions 10 to 40cm across and<br />

5cm high, in three patches in the Veleta (<strong>of</strong> 25, 900 and 1500<br />

square meters respectively) and is abundant over an area <strong>of</strong> 8ha<br />

in the Puerto del Lobo.<br />

The female lays eggs singly inside the leaf rosettes <strong>of</strong> the<br />

foodplant. The first instars feed on the parenchyma <strong>of</strong> the<br />

needle-like leaves and are <strong>of</strong> a purple colour resembling the<br />

dead leaves <strong>of</strong> the plant, among which they usually rest.<br />

Overwintering takes place inside the plant cushions at the third


Hahitat <strong>of</strong> A. zullichi: S. Juan, Sra Nevada, 2760 m, May 1987 (photo by M.L. Munguira).<br />

instar. The last two instars feed on flowers, especially on the<br />

corolla and developing fruits. Their colour is green with yellow,<br />

black and white markings, making them difficult to see in their<br />

colourful environment. They reach their full grown condition<br />

(fifth instar) at the end <strong>of</strong> May after 10 months at the larval<br />

stage. The species is never associated with ants. This fact was<br />

reported by Chapman (1911) for A. glandon that, like A.<br />

zullichi, lacks the dorsal nectary organ found in other lycaenids.<br />

Pupation takes place under stones and adults emerge a month<br />

later (mid-July).<br />

Threats: Development <strong>of</strong> tourist resorts clearly threatens the<br />

survival <strong>of</strong> the Veleta colony: redevelopment <strong>of</strong> the existing ski<br />

station could easily cause the extinction <strong>of</strong> this population.<br />

While the other populations are not threatened by this particular<br />

development they are unprotected at the moment, making<br />

future developments in their areas possible. Collecting could<br />

probably damage the Veleta colony, whose very low population<br />

numbers make it sensitive to any aggression. Large scale<br />

collecting should be banned, because the low total population<br />

numbers <strong>of</strong> the species make any reduction in numbers dangerous<br />

for its future.<br />

<strong>Conservation</strong>: The declaration <strong>of</strong> Sierra Nevada as a Man and<br />

Biosphere Reserve and Natural Park has proved to be ineffective<br />

in protecting the Veleta area from tourist developments. We<br />

79<br />

suggest that the area be declared a National Park. This would<br />

conserve not only its five endemic butterflies (Munguira and<br />

Martin 1989), but also the unusual richness <strong>of</strong> exclusive insects<br />

and plants <strong>of</strong> the area.<br />

The distribution <strong>of</strong> the food plant should be carefully<br />

mapped, in order to identify other possible areas where the<br />

butterfly might be found or introduced if necessary.<br />

Access to the species' habitat should be limited. Particularly,<br />

the construction <strong>of</strong> new roads in the proximity <strong>of</strong> the main<br />

population should be controlled if at all permitted.<br />

Other management practices for habitat improvement are<br />

not needed because <strong>of</strong> the climatic character <strong>of</strong> the plant<br />

communities in which it lives. Therefore the correct policy for<br />

the species conservation is to reduce impacts to a minimum, and<br />

leave the habitat as undisturbed as possible.<br />

References<br />

CHAPMAN, T. A. 1911. On the early stages <strong>of</strong> Latiorina (Lycaena) orbitulus,<br />

an amyrmecophilous Plebeiid Blue butterfly. Trans, ent. Soc. London,<br />

1911: 148–159.<br />

FERNANDEZ-RUBIO, F. 1970. Redescubrimiento de una rara mariposa en<br />

Sierra Nevada. Nota sobre la captura del Lycaenido: Plebejus glandon<br />

zullichi Hemming, 1933 (=nevadensis Zullich y Reisser, 1928). Arch.<br />

Inst. Aclimatacion Almeria 15, 161–167.<br />

HEATH, J. 1981. Threatened Rhopalocera (butterflies) in Europe. Council <strong>of</strong><br />

Europe, Strasbourg.


KUDRNA, O. 1986. <strong>Butterflies</strong> <strong>of</strong> Europe. 8. Aspects <strong>of</strong> the <strong>Conservation</strong> <strong>of</strong><br />

<strong>Butterflies</strong> in Europe. Aula Verlag, Wiesbaden.<br />

MUNGUIRA, M.L. 1989. Biologia y biogeografia de los licenidos ibericos en<br />

peligro de extincion (Lepidoptera, <strong>Lycaenidae</strong>). Ediciones Univ.<br />

Autonoma de Madrid, Madrid.<br />

MUNGUIRA, M.L. and MARTIN, J. 1989. <strong>Biology</strong> and conservation <strong>of</strong> the<br />

80<br />

endangered lycaenid species <strong>of</strong> Sierra Nevada, Spain. Nota lepid., 12<br />

(suppl. 1): 16–18.<br />

VIEDMA, M.G. and GOMEZ, M.R. 1976. Libro Rojo de los lepidopteros<br />

ibericos. ICONA, Madrid.<br />

VIEDMA, M.G. and GOMEZ, M.R. 1985. Revision del Libro Rojo de los<br />

lepidopteros ibericos, ICONA, Monografias no. 42, Madrid.


The Large Copper (Dutch – Grote Vuurvlinder), Lycaena dispar<br />

Haworth<br />

Area: Europe to eastern Asia.<br />

E. DUFFEY<br />

Cergne House, Church Street, Wadenhoe, Peterborough PE8 5ST, U.K.<br />

Status and <strong>Conservation</strong> Interest: Status – L. d. dispar:<br />

extinct; L. d. batava: rare; L. d. rutila: not threatened at present,<br />

(Red List).<br />

A widely distributed wetland species with several described<br />

subspecies. It was first recorded in 1795 in the Huntingdonshire<br />

fens, England. This distinctive and very local subspecies, L. d.<br />

dispar, became extinct in about 1851, at least partly due to<br />

excessive collecting <strong>of</strong> the very locally distributed larvae (Duffey<br />

1968). The Dutch race, L. d. batava, which is very similar to L.<br />

d. dispar, was introduced to Woodwalton Fen Nature Reserve,<br />

England, in 1927 and still survives there.<br />

Taxonomy and Description: L. dispar was first described by<br />

Haworth (1803) and the much more widely distributed L. d.<br />

rutila by Werneburg in 1864. The Dutch race, L. d. batava, was<br />

discovered in 1915 and described by Oberthür in 1923. However,<br />

Higgins and Hargreaves (1983) combine the extinct British<br />

race with the Dutch race under the name L. d. dispar, presumably<br />

because these two races are very difficult to distinguish unless<br />

a series <strong>of</strong> each is available. L. d. dispar and L. d. batava are<br />

generally larger and have more brilliant colours than L. d.<br />

rutila. (See Bink 1970 for further discussion about the<br />

subspecies.)<br />

In the following account the generally used name for the<br />

Dutch population, L. d. batava, is retained.<br />

Distribution: The extinct British population appeared to be<br />

confined to a few fen areas in Huntingdonshire, Cambridgeshire<br />

and Norfolk, although specimens were occasionally taken<br />

elsewhere. The Dutch population <strong>of</strong> L. d. batava is confined to<br />

a few localities in the provinces <strong>of</strong> Friesland and Overijssel<br />

(Bink 1972 and pers. comm. 1991) and has declined in recent<br />

years so that very few colonies survived in 1991. L. d. rutila is<br />

widely but locally distributed in Europe (eastern Germany,<br />

Poland, Baltic countries, Hungary and Russia) but is said to be<br />

declining in the west <strong>of</strong> its range due to drainage and reclamation<br />

<strong>of</strong> wetlands (Higgins and Hargreaves 1983).<br />

81<br />

Population Size: The Dutch population <strong>of</strong> L. dispar has declined<br />

markedly in recent years and only one strong colony is known,<br />

although small numbers occur elsewhere (F.A. Bink, pers.<br />

comm.). The decline is said to be due to lack <strong>of</strong> management <strong>of</strong><br />

the habitat which becomes unfavourable as succession proceeds.<br />

If this situation does not improve, the Dutch population may<br />

become seriously endangered.<br />

The introduced L. d. batava in Britain is very insecure as it<br />

is confined to only about 30ha <strong>of</strong> about 214ha at Woodwalton<br />

Fen. It has been shown that it is vulnerable to excessive summer<br />

flooding (Duffey 1977) and that without protection for the<br />

larvae the population gradually declines to extinction over a<br />

number <strong>of</strong> years. This population has survived since 1927 either<br />

by protecting a large proportion <strong>of</strong> the larvae from predators by<br />

rearing in muslin cages, or else by maintaining a captive<br />

population so that reintroductions can be made when numbers<br />

become dangerously low.<br />

Although the L. d. rutila population has declined in recent<br />

years, it is still widely distributed and not thought to be in<br />

danger.<br />

Habitat and Ecology: Bink (1970,1972,1986) has shown that<br />

L. d. batava in the Netherlands occurs in an area <strong>of</strong> overgrown<br />

peat cuttings where the female lays her eggs on the great water<br />

dock Rumex hydrolapathum in open reed or sedge beds. At<br />

Woodwalton Fen in England it has been shown (Duffey 1977)<br />

that when the foodplants become hidden in taller vegetation the<br />

ovipositing females <strong>of</strong>ten fail to find them. The preferred type<br />

<strong>of</strong> Rumex hydrolapathum is <strong>of</strong>ten <strong>of</strong> moderate size, and large<br />

plants are avoided, especially if growing by open water.<br />

Nevertheless, in years when the butterflies are numerous, eggs<br />

may be laid on all size groups. L. d. batava is single-brooded but<br />

may produce a second brood in warm summers. L. d. rutila is<br />

normally double-brooded. A large female L. d. batava is capable<br />

<strong>of</strong> producing up to 700 eggs but under natural conditions in the<br />

field, production averages 60 per female (Bink 1986) and, at<br />

Woodwalton Fen, 114 per female (Duffey 1968). Bink (1986)<br />

has shown that L. d. batava reaches its highest pupal weight on<br />

host plants growing at pH 5.5–6.5. Below pH 5.5 the foodplants<br />

suffer from acidity stress, resulting in a reduction in the protein<br />

content <strong>of</strong> the leaves. The insects reared on such plants are


smaller and lay fewer eggs. Nevertheless, there is no convincing<br />

evidence that ovipositing females select plants <strong>of</strong> the best<br />

quality.<br />

Eggs are laid singly or in lines, usually along a leaf midrib,<br />

on the underside. However, scattered single eggs are not<br />

infrequent on the upper side <strong>of</strong> the leaves <strong>of</strong> the foodplant. The<br />

larvae hatch after 7–10 days, and graze the leaf surface, forming<br />

'windows'. They hibernate as third-instar larvae in dead leaves<br />

around the base <strong>of</strong> the water dock plants. During hibernation the<br />

larvae are unaffected by winter flooding. In the spring, depending<br />

on weather and the regrowth <strong>of</strong> the docks, the larvae emerge in<br />

late April or early May to feed until they pupate about mid-June.<br />

In July the adults are on the wing, with males usually being the<br />

first to emerge. When a second brood is recorded in L. d. batava<br />

the adults are smaller than the first brood and fewer eggs are<br />

produced. The double broods <strong>of</strong> L. d. rutila have flight periods<br />

<strong>of</strong> May-June and August-September.<br />

Threats: The decline <strong>of</strong> L. dispar, especially in central and<br />

western Europe, is due to the loss <strong>of</strong> wetland habitats where the<br />

preferred foodplant grows. In the Netherlands Bink (1986) has<br />

shown that succession in the fens leads to oligotrophic conditions<br />

which reduces growth and nutrient quality <strong>of</strong> the foodplant.<br />

Fenland reserves in the Netherlands require effective<br />

management to preserve the best conditions for L. dispar. In<br />

England the species is at risk because it occurs at only one<br />

locality and a captive population has to be maintained as an<br />

insurance against extinction.<br />

<strong>Conservation</strong>: Vigorous efforts to protect wetlands are being<br />

made throughout Europe by the International Union for the<br />

<strong>Conservation</strong> <strong>of</strong> Nature and the International Council for the<br />

Protection <strong>of</strong> Birds, but not specifically for invertebrates. The<br />

British population <strong>of</strong> L. d. batava is the responsibility <strong>of</strong> the<br />

Nature Conservancy Council for England (English Nature) and<br />

although a permanent wild population cannot survive in the<br />

82<br />

small area available without larval protection or reintroductions,<br />

no serious attempt has yet been made to assess whether other<br />

suitable areas can be found in the East Anglian wetlands. The<br />

most secure population in the Netherlands is in the Weerribben<br />

fen area. This region could provide a nucleus for re-establishment<br />

in neighbouring fens providing they are maintained in a<br />

favourable condition. There is no doubt that the best prospects<br />

for L. d. batava are in the Netherlands, if the appropriate<br />

authorities could be persuaded to manage suitable fenland areas<br />

more effectively.<br />

Throughout most <strong>of</strong> continental Europe where conservation<br />

work is effective, the main emphasis is on habitat protection for<br />

plants and for vertebrates. More attention should be given to<br />

invertebrates, particularly declining populations such as L. d.<br />

batava and L. d. rutila. The former may soon be endangered and<br />

although the latter is not threatened at present, it may become<br />

so in the future.<br />

References<br />

BINK, F.A. 1970. A review <strong>of</strong> the introductions <strong>of</strong> Thersamonia dispar Haw.<br />

(Lep., <strong>Lycaenidae</strong>) and the speciation problem. Ent. her. Amst. 30:<br />

179–83.<br />

BINK, F.A. 1972. Het onderzoek naar de grote vuurvlinder (Lycaena dispar<br />

batava (Oberthur)) in Nederland (Lep., <strong>Lycaenidae</strong>). Ent. Ber. Amst. 32:<br />

225–39.<br />

BINK, F.A. 1986. Acid stress in Rumex hydrolapathum (Polygonaceae) and<br />

its influence on the phytophage Lycaena dispar (Lepidoptera: <strong>Lycaenidae</strong>).<br />

Oecologia, Berl. 70: 447–451.<br />

DUFFEY, E. 1968. Ecological studies on the large copper butterfly Lycaena<br />

dispar Haw. batavus Obth. at Woodwalton Fen National Nature Reserve,<br />

Huntingdonshire. J. appl. Ecol. 5: 69–96.<br />

DUFFEY, E. 1977. The re-establishment <strong>of</strong> the large copper butterfly Lycaena<br />

dispar batava Obth. on Woodwalton Fen National Nature Reserve,<br />

Cambridgeshire, England, 1969–73. Biol. Conserv. 12: 143–258.<br />

HIGGINS, L. and HARGREAVES, B. 1983. The <strong>Butterflies</strong> <strong>of</strong> Britain and<br />

Europe. Collins, London.


Country: England.<br />

The Adonis Blue, Lysandra bellargus Rottemburg<br />

Status and <strong>Conservation</strong> Interest: Status – locally extinct in<br />

Britain, some rapid decline <strong>of</strong> other colonies: threatened.<br />

Intensive surveys <strong>of</strong> Lysandra (or Polyommatus) bellargus<br />

in the early 1970s revealed that this local species had declined<br />

substantially, and had become rare over much <strong>of</strong> southern<br />

England, the northern fringe <strong>of</strong> its European range. Rapid<br />

losses occurred in the 1950s and the late 1970s and the butterfly<br />

had become extinct on many sites. These included some which<br />

continued to support the Chalkhill Blue, Lysandra coridon<br />

(Poda). In some cases the sites had been destroyed or the<br />

foodplant eliminated by agricultural practices, but disappearance<br />

also from areas <strong>of</strong> unimproved farmland implied that other<br />

effects – perhaps related to grazing regimes – might be involved.<br />

Cool weather was also suggested to be a factor inducing<br />

decline. A study <strong>of</strong> the ecology <strong>of</strong> the species (Thomas 1983)<br />

revealed some unexpected subtleties relevant to the conservation<br />

and management <strong>of</strong> open grassland species, and <strong>of</strong> taxa in<br />

marginally suitable climatic regimes.<br />

Distribution: L. bellargus occurs over much <strong>of</strong> Europe, where<br />

populations have generally not declined as conspicuously as in<br />

Britain. In England, it is confined to calcareous grassland in the<br />

south.<br />

Population Size: This species forms discrete colonies which<br />

commonly contain from about 150–850 individuals. Most<br />

adults do not stray far from the colonies and the populations are<br />

effectively closed; many are isolated from their nearest neighbour<br />

colony by tens <strong>of</strong> kilometres, and no interchange is likely to<br />

occur between colonies even a kilometre or so apart.<br />

Population size within a colony can vary greatly. Heath et<br />

al. (1984) note one Dorset population increasing from fewer<br />

than 50 adults to more than 60,000 between 1977 and 1982.<br />

Such variations provide evidence <strong>of</strong> resilience <strong>of</strong> populations to<br />

extinction (Morris and Thomas 1989) but can also reveal<br />

likelihood <strong>of</strong> extinction: one colony declined from 3400 adults<br />

to extinction in only three years. In the past, L. bellargus has<br />

been recorded from all calcareous formations in southern England.<br />

Extinctions include several colonies in nature reserves.<br />

83<br />

Habitat and Ecology: L. bellargus has two generations each<br />

year. Eggs laid in late August or September hatch into<br />

overwintering caterpillars which mature to adults by around<br />

late May to early June. Offspring <strong>of</strong> this spring generation<br />

develop more rapidly to reach the adult stage in only around 2–3<br />

months. Eggs are laid singly on the foodplant foliage. Larvae<br />

are day-feeders, and all stages from the second instar onward<br />

are tended by ants, mainly Myrmica sabuleti and Lasius alienus.<br />

Larvae and pupae are <strong>of</strong>ten buried by ants, which continue to<br />

tend them.<br />

Larvae feed only on one foodplant species, the horseshoe<br />

vetch Hippocrepis comosa, which occurs much further north in<br />

England than the butterfly does, and is still present in many<br />

southern areas from where L. bellargus has disappeared.<br />

Nearly all populations occur on steep south-facing slopes,<br />

mainly on closely cropped, unimproved pasture. Females prefer<br />

to oviposit in short turf and in sheltered sun-spots. In sites<br />

where there is sward <strong>of</strong> varying heights, oviposition is restricted<br />

almost entirely to short (1–4 cm) vetch, areas which (because <strong>of</strong><br />

high insolation) are both warm and support numerous ants.<br />

Threats: Colony extinction has been due mainly to habitat<br />

change. About one-third <strong>of</strong> colonies were lost because <strong>of</strong> loss<br />

<strong>of</strong> Hippocrepis due to ploughing or 'agricultural improvement'.<br />

However, Hippocrepis persists abundantly on some other sites,<br />

and grazing incidence and intensity are important factors in the<br />

butterfly's well being. Closely cropped sites were very suitable,<br />

although very heavy grazing is harmful. Some major extinctions<br />

coincided with the onset <strong>of</strong> myxomatosis in the 1950s, which<br />

resulted in massive loss <strong>of</strong> rabbits and a resultant decline in<br />

grazing intensity on much chalk grassland. Reduction in grazing<br />

intensity appeared to be a major factor leading to extinction,<br />

and many surviving colonies are on ground grazed by cattle.<br />

'Improved' or lightly grazed sites have only small populations.<br />

However, it is not pr<strong>of</strong>itable to graze unimproved pasture<br />

closely, and some hillsides have been abandoned or are grazed<br />

very irregularly – factors likely to lead to a further decline <strong>of</strong> L.<br />

bellargus on such sites. Cessation <strong>of</strong> grazing can lead to<br />

development <strong>of</strong> coarse grasses and 'choking out' <strong>of</strong> Hippocrepis.<br />

The butterfly's close association with short swards is evident


from its oviposition behaviour, but the reasons behind this are<br />

not clear. L. bellargus is common in tall pasture in parts <strong>of</strong><br />

Europe, for example, and dependence on hotter areas on the<br />

fringe <strong>of</strong> its range might also be a factor influencing site<br />

suitability. Warmth might be important both for the butterfly<br />

itself and for its influence on ants, so that details <strong>of</strong> their<br />

association with L. bellargus' early stages might be very subtle.<br />

Pasture improvement by drilling, herbicides and fertilisers<br />

remains a threat. Until recently, steeper hillsides have not been<br />

ploughed, but some are now cultivated, despite their very thin<br />

soil.<br />

<strong>Conservation</strong>: The above ecological observations (Thomas<br />

1983) emphasise that there is little alternative for practical<br />

conservation <strong>of</strong> the Adonis Blue but to manage sites actively for<br />

its specialised ecological requirements – either on nature reserves<br />

or on commercial farmland, in which case subsidy agreements<br />

may be necessary to ensure site security. Merely reserving<br />

habitat <strong>of</strong> L. bellargus is not sufficient, although reserves are<br />

clearly recommended as a basis for management regimes. For<br />

new reserves, preference may be accorded to ones which<br />

support more than one colony. Management, probably involving<br />

rotational grazing or mowing, must seek to ensure the availability<br />

<strong>of</strong> short sward on south-facing slopes, and that habitats are not<br />

overly fragmented or have barriers (such as valleys or areas <strong>of</strong><br />

tall scrub) imposed between them. Much Hippocrepis has been<br />

converted into a form suitable for L. bellargus during the last<br />

decade by increased rabbit and stock grazing and, if accessible<br />

to a founder population, some such sites have been colonised<br />

successfully.<br />

84<br />

Thomas (1983) also suggested that the carrying capacity <strong>of</strong><br />

many sites for L. bellargus could be increased by creating more<br />

south-facing 'sun-spots', perhaps by using explosives or earthmoving<br />

equipment, and that such methods could be used to<br />

construct sites in previously unsuitable areas.<br />

Another consideration is to introduce L. bellargus to new<br />

areas, or to sites from which it had earlier disappeared, once<br />

these have been rendered suitable again by management. Thus,<br />

L. bellargus was re-introduced in 1981 to Old Winchester Hill<br />

National Nature Reserve, where it had become extinct in the<br />

1950s (Thomas 1989). It increased rapidly in numbers and was<br />

still present after 16 generations (Thomas 1991). Only the<br />

shortest Hippocrepis were utilised, and the breeding sites<br />

circulated in pattern with successive paddocks being grazed<br />

heavily in rotation.<br />

References<br />

HEATH, J., POLLARD, E. and THOMAS, J.A. 1984. Alias <strong>of</strong> <strong>Butterflies</strong> in<br />

Britain and Ireland. Viking Books, Harmondsworth.<br />

MORRIS, M.G. and THOMAS, J.A. 1989. Reestablishment <strong>of</strong> insect<br />

populations, with special reference to butterflies, pp. 22–36 In Emmet,<br />

A.M. and Heath, J. (Eds) The Moths and <strong>Butterflies</strong> <strong>of</strong> Great Britain and<br />

Ireland Vol. 7, Part 1. Harley Books, Great Horkesley.<br />

THOMAS, J.A. 1983. The ecology and conservation <strong>of</strong> Lysandra bellargus<br />

(Lepidoptera: <strong>Lycaenidae</strong>) in Britain. J. appl. Ecol. 20: 59–83.<br />

THOMAS, J.A. 1989. Ecological lessons from the re-introduction <strong>of</strong><br />

Lepidoptera. Entomologist 108: 56–68.<br />

THOMAS, J.A. 1991. Rare species conservation: case studies <strong>of</strong> European<br />

butterflies. pp. 149–197 In: Spellerberg, I.F., Goldsmith, F.B. and Morris,<br />

M.G. (Eds) The Scientific Management <strong>of</strong> Temperate communities for<br />

<strong>Conservation</strong>. Blackwell, Oxford.


Area: England, France, elsewhere in Europe.<br />

Status and <strong>Conservation</strong> Interest: Status – M. arion, M.<br />

alcon, M. teleius: vulnerable (Wells et al. 1983); M. nausithous:<br />

endangered (Wells et al. 1983); M. teleius: endangered (Red<br />

List).<br />

Large Blues were noted by Wells et al. (1983) as 'some <strong>of</strong><br />

the most rapidly declining butterflies in Europe, and probably<br />

in Asia too'. All are threatened with extinction in Europe,<br />

because <strong>of</strong> land use changes (Elmes and Thomas 1992). The<br />

Large Blue, M. arion (L.), became extinct in Britain in 1979<br />

despite valiant long-term efforts to save it, and is extinct also in<br />

the Netherlands, Belgium and parts <strong>of</strong> northern France. The<br />

Alcon Large Blue, M. alcon Denis & Schiffermueller, is also<br />

extinct in many former European localities. The Scarce Large<br />

Blue, M. teleius Bergstrasser, is apparently declining throughout<br />

its European and east Palaearctic range, and is extinct in<br />

Belgium and the Netherlands. The Dusky Large Blue, M.<br />

nausithous Bergstrasser, is also undergoing local extinctions.<br />

Elmes and Thomas (1992) assessed the five European species<br />

as 'Endangered'. Thomas (1984) referred to M. teleius and M.<br />

nausithous as 'among the world's rarest butterflies' and it has<br />

been recommended that research into Maculinea biology be<br />

given top priority in butterfly conservation programmes (Heath<br />

1981). See also Bálint (this volume), for notes on Carpathian<br />

taxa. M. arion in Britain is the best documented conservation<br />

case, and a European subspecies is currently the subject <strong>of</strong><br />

translocations into Britain. This is one <strong>of</strong> very few such<br />

international translocation programmes (see also Duffey, this<br />

volume, on translocation <strong>of</strong> L. dispar). M. arion has been the<br />

target <strong>of</strong> conservation efforts in Britain since the 1920s, latterly<br />

coordinated by a Large Blue Committee, and full-time ecological<br />

work on the British M. arion has been pursued since 1972.<br />

Substantial biological information relevant to conservation <strong>of</strong><br />

some other species has also accumulated (Thomas 1984; Elmes<br />

and Thomas 1992).<br />

Maculinea species are protected legally in Britain, Belgium<br />

and France.<br />

Taxonomy and Description: The British form <strong>of</strong> M. arion was<br />

subspecies eutyphron (Fruhstorfer), and recent successful<br />

Large Blues, Maculinea spp.<br />

85<br />

translocation attempts (Thomas 1989) involve the Swedish M.<br />

a. arion. Rebel's Large Blue,M. rebeli Hirschke, has sometimes<br />

been treated as a subspecies <strong>of</strong> M. alcon, but is distinct<br />

biologically (Elmes and Thomas 1987).<br />

Distribution: Maculinea is Palaearctic. M. arion occurs from<br />

western Europe to southern Siberia, Armenia, Mongolia and<br />

China; M. teleius occurs from Spain to China and parts <strong>of</strong> Japan;<br />

M. nausithous is confined to Europe. The Greater Large Blue,<br />

M. arionides Staudinger, occurs only in China and Japan and its<br />

status is unclear. Wells et al. (1983) list it as 'Vulnerable' but<br />

surveys have not been undertaken in the alpine forest regions it<br />

frequents.<br />

Because several species have been studied in detail they are<br />

treated separately below. Much recent work and synthesis on<br />

the conservation needs <strong>of</strong> Maculinea in western Europe is<br />

included in Elmes and Thomas (1992).<br />

(i) M. arion<br />

Population Size: About 90 sites for M. arion were known in the<br />

southern half <strong>of</strong> England, mainly concentrated in six areas. The<br />

colonies were all circumscribed and most comprised a few tens<br />

to a few hundred adults: the largest probably contained up to<br />

2000–5000 adults in their 'best' years (Thomas and Emmet<br />

1989). In general, colonies were isolated from each other and<br />

closed, as adult dispersal ability is poor.<br />

Extinction <strong>of</strong> the colonies in the six main English areas<br />

occurred as follows: colonies in Northamptonshire died out<br />

around 1860; the last definite record from south Devon was in<br />

1906; colonies in Somerset survived until the late 1950s;<br />

periodic declines in the Cotswolds culminated in the last known<br />

colony disappearing in 1960–1964; the last colony from the<br />

Atlantic coast <strong>of</strong> Devon and Cornwall died out in 1973; and<br />

those in Dartmoor disappeared in the 1970s. Only two sites<br />

remained by 1972, and the butterfly finally became extinct in<br />

Britain in 1979 when the reared female <strong>of</strong>fspring <strong>of</strong> the last<br />

remaining female died before any males emerged which could<br />

mate with them (Thomas 1980).


Habitat and Ecology: M. arion occurs on unimproved grassland<br />

and is ecologically specialised. It is univoltine and adults are<br />

relatively short-lived. Females lay eggs on the flower buds <strong>of</strong><br />

wild thyme, Thymus praecox. The first three caterpillar instars<br />

feed on the thyme flowers and the last (fourth) drops to the<br />

ground and thereafter depends on the attentions <strong>of</strong> Myrmica<br />

ants. As with other Maculinea species, caterpillars are carried<br />

into Myrmica nests, where they feed on ant eggs, larvae and<br />

prepupae; this is its major growth stage. Caterpillars hibernate<br />

and pupate inside ant nests, and the adult M. arion emerges in<br />

late June to mid July.<br />

Any species <strong>of</strong> Myrmica ant will tend the larvae when they<br />

leave the foodplant, but M. arion is essentially specific to<br />

Myrmica sabuleti for successful rearing. Two species <strong>of</strong><br />

Myrmica, sabuleti and scabrinodis, are common where Thymus<br />

grows, but a high density <strong>of</strong> sabuleti with thyme within 2m <strong>of</strong><br />

their nest entrances is needed for the well-being <strong>of</strong> M. arion,<br />

and the size <strong>of</strong> a Large Blue colony was correlated with the<br />

number <strong>of</strong> sabuleti nests present.<br />

The largest colonies in England covered 10–20ha with a few<br />

thousand thyme plants and sabuleti nest densities <strong>of</strong> one every<br />

l–2m 2 . Small colonies occurred on areas <strong>of</strong> less than lha if 60%<br />

<strong>of</strong> the ground was occupied by sabuleti. Thomas (1991) believed<br />

that a 'safe' population <strong>of</strong> 400–1000 M. arion adults could be<br />

supported on lha under ideal conditions: populations with less<br />

than 400 adults (reflecting around 2500 usable sabuleti nests)<br />

might undergo periodic extinctions. Suitable colony sites were<br />

south-facing areas with short-grazed (to about 2cm) turf so the<br />

ground could be sun-baked (Thomas 1980,1991). Sward height<br />

was important: if grazing was removed so that the height<br />

exceeded about 4cm, sabuleti declined substantially, and<br />

scabrinodis became relatively more abundant.<br />

The precise biotope <strong>of</strong> M. arion varies slightly in different<br />

parts <strong>of</strong> Europe, but is always narrow (Thomas 1991).<br />

Threats: Site alienation through improvement for agriculture<br />

(by treatment with herbicides or fertilisers, or more direct<br />

conversion by ploughing or drilling) destroyed about half the<br />

sites and exterminated the colonies on them. The other sites<br />

were mainly abandoned for agriculture, with the resultant<br />

cessation <strong>of</strong> domestic stock grazing, aided in the 1950s by the<br />

spread <strong>of</strong> myxomatosis and removal <strong>of</strong> rabbit grazing. Although<br />

M. sabuleti disappeared rapidly from high sward the thyme<br />

could persist in sward up to 10cm, but it declined in abundance<br />

and few seedlings became established. Several populations <strong>of</strong><br />

M. arion on nature reserves disappeared because <strong>of</strong> lack <strong>of</strong><br />

appreciation <strong>of</strong> the need for grazing management. In hindsight,<br />

with the knowledge now available, it is very likely that extinction<br />

in Britain could have been prevented.<br />

Collecting could have played a role in extinction <strong>of</strong> some<br />

small colonies over the years.<br />

<strong>Conservation</strong>: With knowledge gained in recent years, through<br />

the studies <strong>of</strong> Thomas (1991 and references therein), sites in<br />

Britain suitable for M. arion have been prepared by prescription<br />

grazing and M. sabuleti populations have thereby been increased<br />

86<br />

substantially on several sites as a basis for attempts to reintroduce<br />

M. arion. This commenced in the early 1980s, using Swedish<br />

stock. This was chosen because <strong>of</strong> its phenological suitability:<br />

female butterfly emergence had to coincide with the development<br />

<strong>of</strong> Thyntus flower buds in Britain for oviposition. It was in fact<br />

the only suitable stock available since the Thymus-feeding<br />

races <strong>of</strong> M. arion are all rare in northern Europe. A trial<br />

introduction in 1983 was followed by a major release on one site<br />

in 1986. Seven butterflies emerged in 1984 from the 1983<br />

release and small numbers were present also in 1985 and 1986.<br />

About 200 additional larvae were imported in 1986: some 75<br />

adults emerged in 1987 and 150–200 in 1988. By 1991, it was<br />

estimated (Thomas 1991) that the main site could support<br />

around 600–750 adult butterflies. Introductions have now been<br />

made to other sites and populations will be monitored for<br />

several years. Changes in farming practices (EEC farming<br />

subsidies in the 1980s) and return <strong>of</strong> rabbit grazing have<br />

rendered some <strong>of</strong> the former sites again suitable for M. arion.<br />

The reintroduction has been adjudged successful (Morris and<br />

Thomas 1989; Elmes and Thomas 1992).<br />

Before its extinction in 1979, it had already been anticipated<br />

that reintroduction might be needed, and the work was<br />

coordinated through the 'Joint Committee for the <strong>Conservation</strong><br />

<strong>of</strong> the Large Blue', with several commercial firms providing<br />

sponsorship. The butterfly had already became a familiar emblem<br />

for much other conservation work on British butterflies. It<br />

appeared on stamps and received wide media coverage. One<br />

possible role for a reintroduced colony in due course may be to<br />

serve as a tourist attraction, with controlled access, thus serving<br />

as an important avenue for education on butterfly conservation.<br />

(ii) M. nausithous and M. teleius<br />

Population Size: The small amount <strong>of</strong> published information<br />

suggests that colonies are discrete, closed and generally small<br />

with no more than a few hundred individuals.<br />

Habitat and Ecology. These two species coexist on some sites.<br />

They both breed in marshland, and females oviposit on<br />

flowerbuds <strong>of</strong> the same foodplant, great burnet, Sanguisorba<br />

<strong>of</strong>ficinalis. Because <strong>of</strong> the observed coexistence, it had been<br />

assumed generally that the ecology <strong>of</strong> the two species was very<br />

similar (Thomas 1984) but each has specialised individual<br />

requirements (Elmes and Thomas 1987). As in other species <strong>of</strong><br />

Maculinea, larvae feed on the plant for the first three instars and<br />

then feed on Myrmica brood in ant nests. Larval sizes <strong>of</strong> the two<br />

species differ markedly at the time <strong>of</strong> leaving the plant: an<br />

average larva <strong>of</strong> M. nausithous weighs 1.15mg, and that <strong>of</strong> M.<br />

teleius, 4.32mg. The principle host ant species differ (Thomas<br />

et al. 1989): for M. nausithous it is Myrmica rubra and for M.<br />

teleius, M. scabrinodis, and this segregation helps to explain<br />

why populations <strong>of</strong> Maculinea species have always been<br />

localised in areas where Myrmica ants and foodplants are<br />

abundant. In areas where the butterflies coexist (mainly wetlands<br />

around bogs or in swampy fields) high densities <strong>of</strong> both host<br />

Myrmica species occur.


Threats: Habitat alienation has been attributed to development<br />

and drainage <strong>of</strong> wetlands (Wells et al. 1983); even if the main<br />

reedy areas survive, the drier fringes (where breeding occurs)<br />

may be lost. All known sites for both species in the Rhone<br />

Valley were damaged by reservoir construction in 1981. More<br />

subtle site changes, affecting the abundance and well-being <strong>of</strong><br />

the ants, are also likely to occur: extinction occurs if the density<br />

<strong>of</strong> the particular ant becomes too low and this is occurring<br />

because traditional methods <strong>of</strong> hay and reed cutting are being<br />

abandoned (Thomas 1991). Myrmica scabrinodis is abundant<br />

only in short vegetation, and M. teleius is relatively common<br />

there. M. nausithous gradually replaces it as succession proceeds<br />

and the vegetation becomes taller (4–7 years since<br />

establishment). It may later disappear.<br />

<strong>Conservation</strong>: Reserve establishment to safeguard habitat is a<br />

priority for both species, with management to conserve<br />

plagioclimax conditions. Thomas (1991) notes that high densities<br />

<strong>of</strong> both M. scabrinodis and Sanguisorba can be maintained in<br />

moist hay meadows cut once a year in parts <strong>of</strong> France and<br />

Poland, and there is little doubt that habitats suitable for both<br />

species can be created easily. Translocation may well be a<br />

practical conservation option in the absence <strong>of</strong> natural<br />

colonisation <strong>of</strong> new habitats. Large populations <strong>of</strong> the butterflies<br />

can be supported on small land areas, and vegetation cutting on<br />

a 3-year rotation may be sufficient to maintain site suitability.<br />

(iii) M. alcon and M. rebeli<br />

Population Size: Confusion in the past over the relative status<br />

<strong>of</strong> these two taxa means that much <strong>of</strong> the historical and detailed<br />

distribution <strong>of</strong> each is not wholly clear. Both species occur in<br />

small colonies: many colonies <strong>of</strong> M. alcon contain fewer than<br />

100 individuals.<br />

Habitat and Ecology. The two species can coexist, but M.<br />

rebeli can occur at much higher altitudes than M. alcon and<br />

extends to 1000m in the Swiss alps (Elmes and Thomas 1987).<br />

Eggs <strong>of</strong> both species are laid on Gentiana – those <strong>of</strong> M. alcon<br />

on G. pneumonanthe and G. asclepiadea, and <strong>of</strong> M. rebeli on G.<br />

cruciata and G. germanica. Only large, vigorous plants are<br />

suitable, and the general life history is similar to that <strong>of</strong> other<br />

Maculinea. The main host ant <strong>of</strong> M. alcon is Myrmica ruginodis,<br />

and for M. rebeli, M. schencki.<br />

Threats: Major threats are various forms <strong>of</strong> habitat destruction,<br />

predominantly through agricultural changes, but M. alcon is<br />

87<br />

also threatened to some extent from urbanisation. Elmes and<br />

Thomas (1987) cite specific threats for Swiss populations.<br />

<strong>Conservation</strong>: As for other Maculinea, habitats can be created<br />

by particular mowing or grazing regimes, and reservation with<br />

correct management is needed. A survey <strong>of</strong> the species' range<br />

to detect the best habitats is needed, especially to detail the<br />

distribution <strong>of</strong> M. rebeli in the alps. Recent work by Elmes et al.<br />

(1991a, b) has led to greater understanding <strong>of</strong> the interaction<br />

between the M. rebeli caterpillar and its ant hosts.<br />

References<br />

ELMES, G.W. and THOMAS, J.A. 1987. Le genre Maculinea. pp. 354–368<br />

In: Les Papillons de Jour et leurs Biotopes. Ligue Suisse pour la Protection<br />

de la Nature, Basle.<br />

ELMES, G.W. and THOMAS, J.A. 1992. Complexity <strong>of</strong> species conservation<br />

in managed habitats: interaction between Maculinea butterflies and their<br />

ant hosts. Biodiv. and Conserv. 1: 155–169.<br />

ELMES, G.W., THOMAS, J.A. and WARDLAW, J.C. 1991(a). Larvae <strong>of</strong><br />

Maculinea rebeli, a large-blue butterfly, and their Myrmica host ants:<br />

wild adoption and behaviour in ant-nests. J. Zool. 223: 447–460.<br />

ELMES, G.W., WARDLAW, J.C. and THOMAS, J.A. 1991(b). Larvae <strong>of</strong><br />

Maculinea rebeli, a large-blue butterfly and their Myrmica host ants:<br />

patterns <strong>of</strong> caterpillar growth and survival. J. Zool. 224: 79–92.<br />

HEATH, J. 1981. Threatened Rhopalocera in Europe. Council <strong>of</strong> Europe,<br />

Strasbourg.<br />

MORRIS, M.G. and THOMAS, J.A. 1989. Re-establishment <strong>of</strong> insect<br />

populations, with special reference to butterflies. In: Emmet, A.M. and<br />

Heath, J. (Eds) The Moths and <strong>Butterflies</strong> <strong>of</strong> Great Britain and Ireland.<br />

Vol. 7, Part 1. Harley Books, Great Horkesley, pp. 22–36.<br />

THOMAS, J.A. 1980. Why did the large blue become extinct in Britain? Oryx<br />

15: 243–247.<br />

THOMAS, J.A. 1984. The behaviour and habitat requirements <strong>of</strong> Maculinea<br />

nausithous (the dusky large blue butterfly) and M. teleius (the scarce large<br />

blue) in France. Biol. Conserv. 28: 325–347.<br />

THOMAS, J.A. 1989. The return <strong>of</strong> the Large Blue butterfly. Brit. Wildlife 1:<br />

2–13.<br />

THOMAS, J.A. 1991. Rare species conservation: case studies <strong>of</strong> European<br />

butterflies. In: Spellerberg, I.F., Goldsmith, F.B. and Morris, M.G. (Eds)<br />

The Scientific Management <strong>of</strong> Temperate Communities for <strong>Conservation</strong>,<br />

Blackwell, Oxford, pp. 149–197.<br />

THOMAS, J. A., ELMES, G. W., WARDLAW, J.C. and WOYCIECHOWSKI,<br />

M. 1989. Host specificity among Maculinea butterflies in Myrmica ant<br />

nests. Oecologia 79: 452–457.<br />

THOMAS, J.A. and EMMET, A.M. 1989. Maculinea. In: Emmet, A.M. and<br />

Heath, J. (Eds) The Moths and <strong>Butterflies</strong> <strong>of</strong> Great Britain and Ireland.<br />

Vol. 7, Part 1. Harley Books, Great Horkesley, pp 171–175.<br />

WELLS, S.M., PYLE, R.M. and COLLINS, N.M. 1983. The <strong>IUCN</strong> Invertebrate<br />

Red Data Book. <strong>IUCN</strong>, Gland.


Polyommatus humedasae (Toso & Balletto)<br />

E. BALLETTO<br />

Dipàrtimento di Biologia Animale, Università di Torino, V. Accademia Albertina 17, Torino, Italy –10123<br />

Country: Italy (northwest).<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

This is one <strong>of</strong> the most well known, endemic species living<br />

in a very rare and endangered type <strong>of</strong> habitat, known to be<br />

occupied by a number <strong>of</strong> other restricted insects (Magistretti<br />

and Ruffo 1959, 1960) and plants (Peyronnel 1964). Even<br />

though one or two additional biotopes will probably be<br />

discovered in the future it is doubtful whether those populations<br />

will be as well represented as they are in the species' type<br />

locality.<br />

Taxonomy and Description: Polyommatus humedasae is a<br />

species <strong>of</strong> the subgtnus Agrodiaetus that has been described in<br />

comparatively recent times (Toso and Balletto 1976). Even<br />

though doubts on its species-level identity were initially<br />

expressed by some authors (Higgins and Riley 1983), these<br />

were finally dispelled (Higgins and Hargreaves 1983) by the<br />

study <strong>of</strong> its haploid chromosome complement (n = 38: Troiano<br />

et al. 1979).<br />

The most closely related species, judging by external<br />

morphology, is the Greek Polyommatus aroaniensis (Brown<br />

1976), an endemic <strong>of</strong> the mountains <strong>of</strong> Peloponnesos (Aroania<br />

Ore, i.e. Mt. Chelmos) and characterised by a lower number in<br />

its haploid chromosome complement (n = 15–16).<br />

Distribution: Polyommatus humedasae lives in a particularly<br />

small area <strong>of</strong> a dozen hectares by Pondel (= Pont d'Ael) in the<br />

Val d'Aosta region (northwest Italy). The average altitude <strong>of</strong><br />

the biotope is 1100m and it lies in the montane vegetational<br />

zone.<br />

Population Size: The population structure is probably closed.<br />

Even though the biotope currently known may not be the only<br />

one where this species lives, there is but scanty evidence that<br />

this is the case, mostly based on a single and old museum<br />

specimen and 'entomological gossip'.<br />

Surveys <strong>of</strong> adult specimens demonstrated a reasonably high<br />

population vigour with an average density <strong>of</strong> 11 specimens/<br />

hectare and at least 110 animals instantaneously present at the<br />

site over a period <strong>of</strong> one month. Although no studies focused on<br />

88<br />

the estimation <strong>of</strong> the total population number are available, it<br />

may be represented by a few thousand adults.<br />

Habitat and Ecology: Polyommatus humedasae lives in xerothermophilous<br />

environments <strong>of</strong> the Festucetalia vallesiacae<br />

(Stipo-Poioncarniolocae) vegetational formations; the detailed<br />

association was never investigated to a sufficient detail at the<br />

phyto-sociological level. The geological substrate is represented<br />

by ophiolitiferous calco-schists <strong>of</strong> a Jurassic age (Balletto et al.<br />

1982). As these schists are very fissured and fragmented, they<br />

play a fundamental role in providing a xeromorphic character<br />

to this biotope, irrespective <strong>of</strong> the orientation, which is to the<br />

northeast.<br />

Adults are <strong>of</strong>ten concentrated on what appear to be the<br />

remnants <strong>of</strong> some now abandoned alfalfa fields, whose flowers<br />

represent a good nectar source for all 'blues' <strong>of</strong> the genus<br />

Polyommatus. No male territorial behaviour has been described<br />

or observed.<br />

Eggs are laid singly on the lower surface <strong>of</strong> the leaves <strong>of</strong><br />

Onobrychys montana Lam. & DC. In a laboratory study in<br />

1982, oviposition started in mid-August and peaked between<br />

August 15–21. Hatching starts in the last half <strong>of</strong> September<br />

(Manino et al. 1987). Newly-hatched larvae, lmm long, feed<br />

for a few weeks and then shelter for hibernation. Hibernation<br />

takes place at the first instar, in the litter at the plant base.<br />

Feeding starts again in mid-April, and the second moult takes<br />

place in the first half <strong>of</strong> May. Moults are carried out within the<br />

litter at the base <strong>of</strong> the foodplant. Each moult takes a few days<br />

to perform. Each <strong>of</strong> the successive instars lasts about 15 days.<br />

Full-grown larvae (mid-June) are green and 15mm long.<br />

Pupation also takes place in the litter and lasts about 20–25<br />

days.<br />

No relationship with ants has, so far, been described.<br />

Adults fly from mid-July to mid-August. The butterfly<br />

community <strong>of</strong> Pondel is particularly rich and includes about 30<br />

other species flying synchronously in the same biotope (Balletto<br />

etal. 1982).<br />

Apart from Medicago sativa, adult nectar sources include<br />

Sedum ochroleucum spp. montanum and Onobrychys montana,<br />

the larval food plant. Actual and potential nectar sources are<br />

very abundant throughout the period when imagines are flying


and nectar does not seem to represent a limiting factor <strong>of</strong><br />

population size.<br />

Threats: Potential threats are <strong>of</strong> a rather varied and contrasting<br />

nature. Since the only biotope known for this species is situated<br />

in the montane ecological zone its climax is represented by<br />

woodland. No adult specimens <strong>of</strong> Polyommatus humedasae,<br />

however, have been observed in the woods on any <strong>of</strong> five<br />

surveys carried out in different years. A few small alfalfa fields<br />

that used to be cultivated in this biotope until about ten years<br />

ago have now been abandoned. It seems most likely, therefore,<br />

that if natural succession was to continue towards the climax<br />

vegetation the biotope would disappear, probably together with<br />

the animal. As with many other butterflies, in fact, this appears<br />

to be an ecotonal species, taking advantage <strong>of</strong> the intermediate<br />

stages <strong>of</strong> the recolonisation <strong>of</strong> sun-exposed landslides. It seems<br />

unlikely, however, that in the present situation this 'blue' will<br />

colonise new biotopes by a natural process.<br />

As with many other endemics, another threat is overcollecting.<br />

Even though the exact location <strong>of</strong> the site was not<br />

divulged in the original description in order to prevent such a<br />

threat, collectors soon discovered the place and many <strong>of</strong> them<br />

can <strong>of</strong>ten be met in a single day on the biotope.<br />

<strong>Conservation</strong>: The biotope lies a few hundred metres from the<br />

outer edge <strong>of</strong> the Parco Nazionale del Gran Paradiso, Italy's<br />

largest protected area and one <strong>of</strong> the most strictly regulated. It<br />

seems therefore obvious that a first measure for the conservation<br />

<strong>of</strong> Polyommatus humedasae would be to extend the northern<br />

boundary <strong>of</strong> the Gran Paradiso National Park to include this<br />

biotope (Balletto and Kudrna 1985). The reason why this step<br />

has not yet been taken is apparently that the general area is<br />

heavily under pressure by the conflicting interests <strong>of</strong> the<br />

inhabitants <strong>of</strong> the valleys <strong>of</strong> Cogne and Aosta, who do not want<br />

to renounce rights for agricultural and chamois-stalking<br />

89<br />

practices. Site management, even though not yet needed, may<br />

become necessary in the future, to interrupt the natural trend <strong>of</strong><br />

the vegetation towards a closed woodland.<br />

References<br />

BALLETTO, E., BARBERIS, G. and TOSO, G.G. 1982. Aspetti dell'ecologia<br />

dei Lepidotteri ropaloceri nei consorzi erbacei delle Alpi italiane. Collana<br />

'Promozione della qualita dell'ambiente': Quaderni sulla 'struttura<br />

delle Zoocenosi terrestri' II.2: 11–96.<br />

BALLETTO, E. and KUDRNA, 0.1985. Some aspects <strong>of</strong> the conservation <strong>of</strong><br />

butterflies in Italy, with recommendations for a future strategy. Boll. Soc.<br />

ent. ital. 117 (1–3): 39–59.<br />

BROWN, J. 1976. Notes regarding previously undescribed European taxa <strong>of</strong><br />

the genera Agrodiaetus Hubner, 1822 and Polyommatus Kluk, 1801.<br />

Entomologist's Gaz. 27: 27–83.<br />

HIGGINS, L.G. and HARGREAVES, B. 1983. The butterflies <strong>of</strong> Britain and<br />

Europe. Collins, London.<br />

HIGGINS, L.G. and RILEY, N.D. 1983. A field guide to the butterflies <strong>of</strong><br />

Britain and Europe, 5th ed., 384pp. Collins, London.<br />

MAGISTRETTI, M. and RUFFO, S. 1959. Primo contributo alla conoscenza<br />

della fauna delle oasi xerotermiche prealpine. Mem. Mus. civico St. nat.<br />

Verona 7: 99–125.<br />

MAGISTRETTI, M. and RUFFO, S. 1960. Secondo contributo alla conoscenza<br />

della fauna delle oasi xerotermiche prealpine. Mem. Mus. civico St. nat.<br />

Verona 8: 223–240.<br />

MANINO, Z., LEIGHEB, G., CAMERON-CURRY, P. and CAMERON-<br />

CURRY, V. 1987. Descrizione degli stadi preimmaginali di Agrodiaetus<br />

humedasaeToso & Balletto, 1976 (Lepidoptera, <strong>Lycaenidae</strong>). Boll. Mus.<br />

regionale Sci. nat. Torino 5(1): 97–101.<br />

PEYRONNEL, B. 1964. Escursione della Societa Botanica Italiana in Val<br />

d'Aosta. Giorn. bot. ital. 71: 183–196.<br />

TOSO, G.G. and BALLETTO, E. 1976. Una nuova specie del genere<br />

Agrodiaetus Hubn. (Lepidoptera: <strong>Lycaenidae</strong>). Annali Mus. civico Sty.<br />

nat. G. Doria, Genova 81: 124–130.<br />

TROIANO, G., BALLETTO, E. and TOSO, G.G. 1979. The karyotype <strong>of</strong><br />

Agrodiaetus humedasae Toso & Balletto (Lepidoptera: <strong>Lycaenidae</strong>).<br />

Boll. Soc. ent. ital. 111(7–8): 141–143.


Polyommatus galloi (Balletto & Toso)<br />

E. BALLETTO<br />

Dipàrtimento di Biologia Animale, Università di Torino, V. Accademia Albertina 17, Torino, Italy — 10123<br />

Country: Italy (south)<br />

Status and <strong>Conservation</strong> Interest: Status – rare.<br />

This is one <strong>of</strong> the few relatively well known species living<br />

in a very rare and endangered type <strong>of</strong> habitat, known to be<br />

inhabited by a number <strong>of</strong> exclusive insects and plants (Gavioli<br />

1936; Avena and Bruno 1975).<br />

Taxonomy and Description: Polyommatus galloi is another<br />

species <strong>of</strong> the subgenus Agrodiaetus Hübner described in<br />

comparatively recent times (Balletto and Toso 1979). Its specieslevel<br />

distinction from Polyommatus ripartii (Freyer) (south<br />

France, northwest Italy) was confirmed by the study <strong>of</strong> the<br />

haploid chromosome complement (n = 66: Troiano 1979,<br />

instead <strong>of</strong> n = 90 as in P. ripartii: Lesse 1960).<br />

The most closely related species, from both external<br />

morphology and haploid chromosome complement, is<br />

apparently Polyommatus demavendi (Pfeiffer), distributed from<br />

Turkey to north Iran and characterised by a slightly larger<br />

number <strong>of</strong> chromosomes (n = 70–71: Lesse 1960).<br />

Distribution: Polyommatus galloi lives over an area <strong>of</strong> several<br />

square kilometres shared between the southern Italian regions<br />

<strong>of</strong> Calabria and Lucania, on Mt Pollino and on the Orsomarso<br />

mountain range. The altitude <strong>of</strong> biotopes inhabited by this<br />

species ranges from 1800 to 2200m.<br />

Population Size: The population structure is relatively closed<br />

in that specimens at high altitudes on Mt Pollino are unlikely to<br />

be able to reach the Orsomarso chain, and vice versa. Each <strong>of</strong><br />

the two main sets <strong>of</strong> biotopes contain an apparently small<br />

number <strong>of</strong> metapopulations. Even though the biotopes currently<br />

known may not be the only ones where this species lives, there<br />

is currently no evidence that this is the case.<br />

Surveys <strong>of</strong> adult specimens demonstrated a reasonably high<br />

population density with an average 6–7 specimens/hectare<br />

simultaneously present at the same site over a period <strong>of</strong> about<br />

one month. On the whole this species may be represented by<br />

several thousand adults/year.<br />

90<br />

Habitat and Ecology: Polyommatus galloi lives in xeromorphic<br />

environments, at the highest elevations reached in the southern<br />

Italian Apennines by the vegetational formations <strong>of</strong> the Alliance<br />

Bromion erecti. The particular Sub-Alliance represented here<br />

(Seslerio-Xerobromenion apennium) is considered somewhat<br />

transitional to the Seslerietalia apenninae present at the high<br />

elevations <strong>of</strong> the central Apennines in otherwise similar<br />

ecological conditions (Avena and Bruno 1975). The geological<br />

substrate is represented by Lower Cretaceous grey limestones,<br />

or by Jurassic calcitutites and limestones (Balletto et al. 1982).<br />

Adults are generally concentrated on the flowerheads <strong>of</strong> the<br />

nectar sources. No male territorial behaviour has been described<br />

or observed. No particular study has been devoted to the<br />

reproductive biology <strong>of</strong> this species but females have been<br />

observed to lay eggs on the lower surface <strong>of</strong> the leaves <strong>of</strong><br />

Onobrychis caputgalli. (L.) Lam. Adults fly from mid-July to<br />

mid-August.<br />

No relationship with ants has, so far, been described or<br />

observed.<br />

Apart from Lavandula angustifolia ssp. angustifolia, a good<br />

nectar source for most xerophilous species <strong>of</strong> Polyommatus,<br />

adults feed on the flowerheads <strong>of</strong> Sedum album, Picris<br />

hieracioides, Echinops ritro, Cirsium afrum and Onobrychis<br />

caputgalli. The latter also represents the larval foodplant.<br />

Actual and potential nectar sources are very abundant throughout<br />

the period when imagines are flying and nectar is not likely to<br />

represent a limiting factor for population size.<br />

Threats: A study conducted in 1977 and again in 1980–81<br />

(Balletto et al. 1982) has shown that this species is particularly<br />

sensitive to the adverse influence <strong>of</strong> grazing. Even though the<br />

effect <strong>of</strong> sheep overgrazing is particularly severe, even slight<br />

grazing by domestic stock can result in considerably diminished<br />

population densities, which soon fall well below 50% <strong>of</strong> normal.<br />

A potential threat is over-collecting although for the time being<br />

this is not an issue.<br />

<strong>Conservation</strong>: The most important thing would be to reverse<br />

the negative effects <strong>of</strong> sheep grazing.


Mt. Pollino is included in a Natural Park, one <strong>of</strong> the largest<br />

in southern Italy, but unfortunately not one having particularly<br />

strict regulations with regard to stock rearing practices (Balletto<br />

and Kudrna 1985). The main reason why, up until now, steps in<br />

this direction have not yet been taken, is that the general area is<br />

under pressure from the conflicting interests <strong>of</strong> the local<br />

shepherds, who do not want to renounce, or even reduce their<br />

grazing rights.<br />

In the Mediterranean area, where summer rainfall is episodic,<br />

overgrazing causes the soil moisture to evaporate very quickly<br />

and in the long run can easily elicit severe transformations <strong>of</strong> the<br />

habitat <strong>of</strong> this species.<br />

Site management, even though not yet needed, may become<br />

necessary on a local basis, to interrupt the natural trend <strong>of</strong> the<br />

vegetation towards a closed woodland.<br />

91<br />

References<br />

AVENA.G. and BRUNO, F. 1975. Lineamenti della vegetazione del Masiccio<br />

del Pollino (Appennino calabro-lucano). Not. Fitosoc. 10: 131–153,<br />

Roma.<br />

BALLETTO, E. and KUDRNA, O.1985. Some aspects <strong>of</strong> the conservation <strong>of</strong><br />

butterflies in Italy, with recommendations for a future strategy. Boll. Soc.<br />

ent. ital. 117(1–3): 39–59.<br />

BALLETTO, E. and TOSO, G.G. 1979. On a new species <strong>of</strong> Agrodiaetus<br />

(<strong>Lycaenidae</strong>) from Southern Italy. Nota lepid. 2(1/2): 13–25.<br />

BALLETTO, E., TOSO, G.G. and BARBERIS, G. 1982. Le comunita di<br />

Lepidotteri ropaloceri nei consorzi erbacei dell'Appennino. Collan<br />

'Promozione della qualita dell'ambiente': Quaderni sulla 'struttura<br />

delle Zoocenosi terrestri' II.l: 77–143.<br />

GAVIOLI, O.1936. Ricerche sulla distribuzione altimetrica della vegetazione<br />

in Italia: 3°. Limiti altimetrici delle formazioni vegetali nel gruppo del<br />

Pollino (Appennino calabro-lucano). N. Giorn. bot. ital., n.s. 43.<br />

LESSE, H. (de) 1960. Les nombres de chromosomes dans le groupe<br />

d'Agrodiaetus ripartii Freyer (Lepidoptera <strong>Lycaenidae</strong>). Revue fr. Ent.<br />

27(3): 240–264.<br />

TROIANO, G. 1979. Karyotype. In: Balletto E., & Toso G.G., On a new<br />

species <strong>of</strong> Agrodiaetus (<strong>Lycaenidae</strong>) from Southern Italy. Nota lepid. 2(1/<br />

2): 13–25.


The Sierra Nevada Blue, Polyommatus golgus (Hübner)<br />

M.L. MUNGUIRA and J. MARTIN<br />

Deparlamento de Biologia (Zoologia), Facultad des Ciencìas, Universidad Autonoma de Madrid, Madrid, Spain<br />

Country: Spain (southeast).<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable (listed<br />

as endangered (Viedma and Gomez 1985) or vulnerable (Heath<br />

1981)).<br />

This species, also known as the Nina de Sierra Nevada, and<br />

Agriades zullichi are flagship representatives <strong>of</strong> a very peculiar<br />

and endangered habitat. They live in an area with 53 endemic<br />

plant species and at least 50 exclusive insects (including three<br />

endangered Orthoptera). Several authors have supported the<br />

idea <strong>of</strong> protecting the area as a National Park (Gómez-Campo<br />

1987; Munguira and Martin 1989a), but part has been developed<br />

as a ski station, creating a conflict between conservation and<br />

development in an area where job creation and investment are<br />

badly needed.<br />

P. golgus is listed in Appendix II <strong>of</strong> the Berne Convention<br />

<strong>of</strong> which Spain is a Contracting Partner.<br />

Taxonomy and Description: Hübner described golgus as a<br />

distinct species, but it was considered by many authors to be a<br />

subspecies <strong>of</strong> P. dorylas (Denis & Schiffermuller) (Agenjo<br />

1947). Lesse (1960) studied the chromosome numbers <strong>of</strong> the<br />

dorylas group and stated that golgus had a lower number (n =<br />

c. 131–134) than the other species. This has been used by other<br />

authors (Gomez and Arroyo 1981) as evidence in favour <strong>of</strong> it<br />

being a true species.<br />

Another closely related species, P. nivescens (Keferstein),<br />

with n = c. 190–191 (one <strong>of</strong> the highest chromosome numbers<br />

in the animal kingdom) is endemic to Spain and also lives in the<br />

Sierra Nevada at lower altitudes; and an ecologically similar<br />

species (P. atlantica (Elwes)) is endemic to the Atlas Mountains<br />

in Morocco.<br />

Distribution: Restricted to four 10x10km UTM squares in the<br />

Sierra Nevada (Granada Province, southeast Spain). It lives at<br />

heights ranging from 2500 to 2900m in the oromediterranean<br />

and crioromediterranean zones (the Mediterranean equivalents<br />

<strong>of</strong> subalpine and alpine zones).<br />

Population Size: Population structure is not closed and small<br />

numbers <strong>of</strong> adults can be seen all around the suitable area.<br />

92<br />

Larval surveys revealed a very low population density, but the<br />

populations occupied a very extensive area. Adult males are<br />

concentrated in wet places where they defend perching sites<br />

against other conspecific males in a behaviour similar to lek<br />

behaviour (Munguira and Martin 1989b). The population studied<br />

probably had several thousand adults, but no accurate estimate<br />

has been made. Records <strong>of</strong> the species from 1838 to 1986 show<br />

it as abundant in the habitat to which it is restricted.<br />

Habitat and Ecology: Present in grassland communities<br />

growing among dwarf junipers (Genisto baeticae-Junipereto<br />

nanae) and at higher altitudes on climax grasslands (Erigeronto<br />

frigidi-Festuceto clementei) growing among schist screes<br />

(Munguira 1989; Munguira and Martin 1989b). Due to the very<br />

harsh weather conditions (with snow cover for nine months) the<br />

plants growing in the area are strong rooted perennials with<br />

aerial parts growing close to the ground. One <strong>of</strong> these plants is<br />

Anthyllis vulneraria arundana, the butterfly's foodplant, which<br />

is also endemic to the area.<br />

P. golgus, in cop., Sierra Nevada, July 1985 (photo by M. Munguira).


Figure 2. Habitat <strong>of</strong> P. golgus, Sierra Nevada, Veleta, September 1986 (M. Munguira).<br />

Eggs are laid singly on the upperside <strong>of</strong> curled leaves <strong>of</strong> the<br />

plant, whose parenchyma is used as food by the caterpillars.<br />

The species overwinters in its third larval instar. Larvae are<br />

regularly tended by Tapinoma nigerrimum ants, that <strong>of</strong>ten have<br />

their nests close to the foodplants. Pupation takes place in June,<br />

after five larval instars, in the ground near the foodplant. Adults<br />

fly in July in a single generation.<br />

Nectar sources include Arenaria tetraguetra, Silene<br />

rupestris, Jasione amethystina and Hieracium pilosella. The<br />

flowers <strong>of</strong> these plants and many others are abundant in the area<br />

during the flight period, and therefore nectar does not seem to<br />

be a limiting factor for the species.<br />

Threats: Tourism-related development is the major threat for<br />

the butterfly. A metalled road crosses one <strong>of</strong> the areas and a ski<br />

station actually exists in part <strong>of</strong> the butterfly's habitat.<br />

Redevelopment <strong>of</strong> the ski station poses the greatest threat,<br />

involving construction <strong>of</strong> new roads and buildings and reshaping<br />

<strong>of</strong> slopes for ski courses. These would have great impact on the<br />

butterfly's habitat and make other further impacts (pollution,<br />

refuse accumulation) more likely.<br />

<strong>Conservation</strong>: The climax character <strong>of</strong> the plant communities<br />

in which the butterfly lives is a great advantage for conservation<br />

93<br />

practice: the only necessary action is to protect the area and try<br />

to reduce impacts to their minimum. The declaration <strong>of</strong> a<br />

National Park in the area seems necessary, at least to protect one<br />

<strong>of</strong> the richest mountain ranges in Europe as far as endemic<br />

plants is concerned (Blanco 1989), as well as the five endemic<br />

butterflies (Parnassius apollo nevadensis Oberthur, Erebia<br />

hispaniaBut\cT,Polyommatusgolgus,Agriadeszullichi,Aricia<br />

morronensis ramburi) and other insects. New developments in<br />

the Monachil and Dilar valleys should be stopped and any jobcreating<br />

alternatives studied so that conservation does not<br />

necessarily involve refusal to develop a depressed area. Sierra<br />

Nevada was declared in 1986 a reserve <strong>of</strong> the Man and Biosphere<br />

(MAB) project, and in 1989 it became a Natural Park, but this<br />

conservation status has not prevented the area from being<br />

damaged.<br />

The inclusion <strong>of</strong> P. golgus in the appendices <strong>of</strong> the Berne<br />

Convention probably assures its protection in the long term but<br />

no short-term actions have been undertaken, and the<br />

developments currently taking place have not been stopped<br />

despite legislation against them.<br />

This is one <strong>of</strong> the most obvious cases in which the importance<br />

<strong>of</strong> education should be stressed. Unlike other National Parks, in<br />

which the conservation interests are geological formations,<br />

remarkable forests or vertebrate faunas, the principal values <strong>of</strong>


Sierra Nevada are its small endemic invertebrates and plants. Its<br />

high altitude and geographical location makes it the southernmost<br />

limit for many northern and alpine plants and animals. Its<br />

populations are therefore <strong>of</strong> key importance in conserving the<br />

genetic diversity <strong>of</strong> these species. It is therefore necessary to<br />

enhance awareness among the general public <strong>of</strong> the great<br />

scientific and conservation interest <strong>of</strong> this apparently unattractive<br />

area.<br />

References<br />

AGENJO, R. 1947. Catálogo ordenador de los lepidópteros de España.<br />

Sexagesima novena familia. Graellsia 5.<br />

BLANCO, E. 1989. Areas y enclaves de interés botánico en España (Flora<br />

silvestre y vegetación). Ecologia 3: 7–21.<br />

G6MEZ, M.R. and ARROYO, M. 1981. Catálogo sistemático de los<br />

94<br />

lepidópteros ibéricos. Ministerio de Agricultura y Pesca, Madrid.<br />

G6MEZ-CAMP0, C. (Ed.) 1987. Libra rojo de especies vegetates amenazadas<br />

de España peninsular e Islas Baleares. ICONA, Madrid.<br />

HEATH, J. 1981. Threatened Rhopalocera (butterflies) in Europe. Council <strong>of</strong><br />

Europe, Strasbourg.<br />

LESSE, H. De. 1960. Spéciation et variation chromosomique chez les<br />

lépidoptères rhopaloceres. Ann. Sci. Nat. Zool. Biol. Anim. 2: 1–223.<br />

MUNGUIRA, M.L. 1989. Biologiaybiogeografla de los licénidos ibéiricos en<br />

peligro de extincién (Lepidoptera, <strong>Lycaenidae</strong>). Ediciones Universidad<br />

Autónoma de Madrid, Madrid.<br />

MUNGUIRA, M.L. and MARTIN, J.1989a. <strong>Biology</strong> and conservation <strong>of</strong> the<br />

endangered lycaenid species <strong>of</strong> Sierra Nevada, Spain. Nota lepid. 12<br />

(suppl. 1): 16–18.<br />

MUNGUIRA, M.L. and MARTIN, J. 1989b. Paralelismo en la biologia de tres<br />

especies taxonómicamente próximas y ecológicamente diferenciadas del<br />

género Lysandra: L. dorylas, L. nivescens y L. golgus (Lepidoptera,<br />

<strong>Lycaenidae</strong>). Ecologia 3: 331–352.<br />

VIEDMA, M.G. and GOMEZ, M.R. 1985. Revisión del Libro Rojo de los<br />

lepidópteros ibéricos. ICONA, Madrid.


Le Faux-Cuivre smaragdin*, Tomares ballus F.<br />

Henri A. DESCIMON<br />

Laboratoire de Systématique evolutive, Université de Provence, 3 place Victor Hugo, 13331 Marseille Cedex 3, France<br />

Country: Western Mediterranean (France).<br />

Status and <strong>Conservation</strong> Interest: Status – not thought to be<br />

threatened at the present time.<br />

This species may be a test one, facing both the impetus <strong>of</strong><br />

economic development in the Mediterranean region, especially<br />

in the coastal belt where its habitats occur, and the abandonment<br />

<strong>of</strong> traditional land occupation, which provides its chief haunts.<br />

Taxonomy and Description: The Genus Tomares Rambur is<br />

limited to the western Palaearctic region. Three <strong>of</strong> the six<br />

species recognized are distributed around the Mediterranean<br />

basin. T. ballus occurs as two subspecies: the nominal one, from<br />

north Africa and southern Spain, and catalonica Sagarra, from<br />

Spanish Catalonia and southeast France. The latter is<br />

characterised by a yellowish-green hue <strong>of</strong> the hindwing underside<br />

while the nominal subspecies is bluish-green.<br />

Distribution: T. ballus is widely distributed in Maghreb and<br />

the southern half <strong>of</strong> Spain (Gomez-Bustillo and Fernandez-<br />

Rubio 1974). In France, it is limited to a nucleus disjunct from<br />

the main area in the littoral region <strong>of</strong> the Var and Alpes-<br />

Maritimes, with some recently discovered colonies in Bouchesdu-Rhone.<br />

Although still widespread in the Var department and<br />

able to colonise available habitats rapidly, the species has<br />

undergone some restriction <strong>of</strong> its geographical range in France.<br />

It no longer exists in the Alpes Maritimes, where it was present<br />

until the 1970s at Cannes and Vallauris. It has also been<br />

eliminated in its classical localities close to Hyeres.<br />

Population Size: In southern Spain, the species is <strong>of</strong>ten very<br />

abundant (Jordano et al. 1990a) and widespread. In France, the<br />

colonies are patchy and unstable, although they appear to be<br />

dense (several hundreds/ha, according to the author's visual<br />

estimations). However, the species displays obvious colonising<br />

abilities, and starts thriving in new available habitats within a<br />

very few years. Episodical colonisation <strong>of</strong> localities outside its<br />

normal breeding range is occasionally reported (de Laever<br />

1954).<br />

* A name contrived recently by G.C. Luquet.<br />

95<br />

Habitat and Ecology. Throughout its area <strong>of</strong> distribution, the<br />

species is confined to open landscapes, either steppe vegetation<br />

formation, or forest clearings <strong>of</strong> a sufficient size, where the<br />

vegetation covering is sparse. In France, it is confined to<br />

calcareous substrates. In this country, there are two main kinds<br />

<strong>of</strong> habitat: forest clearings <strong>of</strong> the live oak-Aleppo pine forest;<br />

and semi-neglected orchards <strong>of</strong> olive trees and vines, where by<br />

far the more abundant colonies are to be found.<br />

Foodplants differ according to the geographical region but<br />

always belong to Fabaceae: in northern Africa, mainly Erophaca<br />

baetica (Powell in Oberthur 1910); in southern SpainAstragalus<br />

lusitanicus (Sierra Morena) and Medicago polymorpha<br />

(Guadalquivir Valley) (Jordano et al. 1990a, 1990b); in France,<br />

either Bonjeana hirsuta (Chapman 1904) or predominantly<br />

Anthyllis tetraphylla, which is also used in northern Morocco<br />

(Descimon and Nel 1986). B. hirsuta is chiefly confined to the<br />

woodland clearings habitat, while A. tetraphylla used to thrive<br />

in cultivated areas where weeding was regular enough to<br />

prevent scrub invasion but sparse enough to allow annuals,<br />

therophytes and hemicryptophytes to grow.<br />

Feeding behaviour appears to be highly opportunistic and<br />

depends upon the compatibility between the vegetative cycle <strong>of</strong><br />

the plant and the developmental stage <strong>of</strong> the butterfly. Fine<br />

adjustment <strong>of</strong> the latter to the constraints imposed by the<br />

various host-plant cycles is evident.<br />

In France, eggs are laid singly or in very small groups, while<br />

in Sierra Morena, they are laid in clumps <strong>of</strong> several tens. The<br />

caterpillars prefer to consume flowers and seedpods, but can<br />

also attack leaves. In France, the duration <strong>of</strong> the caterpillar stage<br />

is 50–60 days, depending on yearly climate. Pupation occurs<br />

under stones present in the habitat. Pupal duration is about 10<br />

months.<br />

Threats: In France, near the seashore, the only cause <strong>of</strong><br />

disappearance <strong>of</strong> T. ballus is the rampant urbanisation <strong>of</strong> the<br />

strip <strong>of</strong> land bordering the crowded beaches <strong>of</strong> the Mediterranean<br />

sea. On the mainland, improved cultivation techniques quickly<br />

cause the disappearance <strong>of</strong>'weeds' such as Anthyllis tetraphylla.<br />

In woodland areas, the dramatically increased frequency <strong>of</strong>


fires (<strong>of</strong> course due to urbanisation) is also a cause <strong>of</strong> the<br />

extinction <strong>of</strong> some colonies, in particular those close to Hyeres.<br />

However, in this case, the return <strong>of</strong> the butterfly following<br />

resprouting <strong>of</strong> its foodplants can be rapid if nearby colonies are<br />

preserved. In all zones, nibbling ('moth-eating' in French!) <strong>of</strong><br />

the countryside by private houses and their gardens specifically<br />

attacks the chief haunt <strong>of</strong> T. ballus: sunny, terraced, sparsely<br />

cultivated landscapes. In the zones not yet struck by building<br />

speculation (and they become scarce, since, under the southern<br />

sun, every piece <strong>of</strong> land is under threat), land abandonment<br />

causes the open therophytic landscapes to be invaded by scrub<br />

and then continuous pine forest with no more suitable habitats<br />

available to T. ballus.<br />

The low yield <strong>of</strong> traditional Mediterranean cultivation<br />

practices does not now allow for continuing a landscape<br />

maintenance regime which, in the past, provided a high floristic<br />

and faunistic diversity.<br />

Collecting does not play any role in the decline <strong>of</strong> the<br />

species: even in the limited areas where overcollecting was<br />

exercised by collectors when the habitats remained ecologically<br />

preserved, a decrease in the species' abundance was scarcely<br />

observed.<br />

<strong>Conservation</strong>: At present, no legal measures for protection are<br />

taken in France. Should such measures be taken, they would<br />

probably be limited to prohibiting collection. Such provisions<br />

prove extremely inefficient – especially in a country <strong>of</strong> Latin<br />

and Mediterranean tradition – and can even be<br />

counterproductive: pr<strong>of</strong>essional entomologists waste much time<br />

in satisfying administrative formalities; honest amateur<br />

entomologists are discouraged and give up butterfly watching;<br />

dishonest collectors ignore the laws completely; and<br />

unscrupulous dealers continue to enjoy the increased prices <strong>of</strong><br />

'black market' specimens.<br />

The problem <strong>of</strong> conservation <strong>of</strong> a large, semi-continuous<br />

and widespread population <strong>of</strong> Tomares ballus in southern<br />

France is closely linked to the general problem <strong>of</strong> overall<br />

96<br />

conservation <strong>of</strong> biological diversity and even <strong>of</strong> human living<br />

quality in this region. Up to the present time, in spite <strong>of</strong> some<br />

pungent attacks through urbanisation <strong>of</strong> the seashore region,<br />

southern France has been relatively safe from the 'economic<br />

development syndrome' which rages in northern Europe and<br />

involves intensified exploitation <strong>of</strong> pr<strong>of</strong>it-earning zones,<br />

abandonment <strong>of</strong> other zones, overurbanisation and general<br />

pollution (Descimon 1990).<br />

At the present time, perhaps the most effective strategy to<br />

conserve T. ballus habitat would be an educational effort<br />

directed towards private landowners in 'moth-eaten'<br />

countryside: such efforts could maintain parts <strong>of</strong> the traditional<br />

Meditteranean landscape, such as olive orchards, with a low<br />

level <strong>of</strong> 'cleaning'.<br />

In Spain, where the species is more abundant, and Maghreb<br />

(Thomas and Mallorie 1985), the problems are less serious but<br />

probably basically the same.<br />

References<br />

DE LAEVER, E. 1954. Tomares ballus Fabr. dans les Basses Alpes. R. fr.<br />

Lépidoptérologie 14: 165.<br />

DESCIMON, H. and NEL, J. 1986. Tomares ballus F. est-il une espéce<br />

vulnérable en France? Alexanor 14: 219–231.<br />

DESCIMON, H. 1990. Pourquoi y a-t-il moins de papillons aujourd'hui?<br />

Insectes 77: 6–10.<br />

GOMEZ-BUSTILLO, M.R. and FERNANDEZ-RUBIO, F. 1974. Mariposas<br />

de la Peninsula iberica. Ropaloceros (II). Ministerio de Agriculture,<br />

Madrid. 258pp.<br />

JORDANO, D., HAEGER, J.F. and RODRIGUEZ, J. 1990a. The effect <strong>of</strong><br />

seed production by Tomares ballus (Lepidoptera: <strong>Lycaenidae</strong>) on<br />

Astragalus lusitanicus (Fabaceae): determinants <strong>of</strong> differences among<br />

patches. Oikos 57: 250–256.<br />

JORDANO, D., HAEGER, J.F. and RODRIGUEZ, J. 1990b. The life-history<br />

<strong>of</strong> Tomares ballus (Fabricius, 1787) (Lepidoptera: <strong>Lycaenidae</strong>): phenology<br />

and host plant use in southern Spain. J. Res. Lepid. 28: 112–122.<br />

THOMAS, CD. and MALLORIE, H.C. 1985. Rarity, species richness and<br />

conservation: butterflies <strong>of</strong> the Atlas mountains in Morocco. Biol. Conserv.<br />

33: 95–117.


Country: U.K.<br />

The Silver-studded Blue, Plebejus argus L.<br />

CD. THOMAS<br />

Centre for Population <strong>Biology</strong>, Imperial College at Silwood Park, Ascot, Berks SL5 7PY, U.K.<br />

Status and <strong>Conservation</strong> Interest: Status – P. a. masseyi:<br />

extinct; P. a. cretaceus: not threatened but few colonies survive;<br />

P. a. caernensis: not threatened; P. a. argus: not threatened but<br />

few colonies survive.<br />

P. argus has virtually disappeared from four-fifths <strong>of</strong> its<br />

former British range (Heath et al. 1984; Thomas and Lewington<br />

1991). Studies have been undertaken in north Wales (CD.<br />

Thomas 1985a,b, 1991; Thomas and Harrison 1992), Devon<br />

(Read 1985), Suffolk (Ravenscr<strong>of</strong>t 1986,1987,1990), Norfolk<br />

(N.Armour-Chelu, pers. comm.), and Dorset (J.A. Thomas<br />

1991). P. argus is regarded as an important indicator <strong>of</strong> vigorous<br />

heathland habitats, and has been severely affected by habitat<br />

fragmentation, and by the cessation <strong>of</strong> traditional management<br />

which maintains the heathland successions required by this<br />

insect. Many <strong>of</strong> the British populations <strong>of</strong> P. argus are already<br />

on reserves, and English Nature, the Countryside Council for<br />

Wales, The Royal Society for the Protection <strong>of</strong> Birds (RSPB),<br />

the National Trust, the County Naturalist Trusts, and Local<br />

Councils actively foster this species.<br />

Taxonomy and Description: P. argus has formed several local<br />

races in Britain, and although these probably do not warrant the<br />

formal status <strong>of</strong> subspecies (CD. Thomas 1985a), they do<br />

provide extra conservation interest. Race masseyi was found on<br />

the mosslands at the southern margin <strong>of</strong> the Lake District, and<br />

had blue females (de Worms 1949). Scottish populations were<br />

similar. Race caernensis is restricted to limestone grasslands in<br />

north Wales. They are very small, and have blue females.<br />

Heathland populations in north Wales look similar to caernensis,<br />

but are slightly larger, and a mossland population in north<br />

Wales is similar but distinctly larger (CD. Thomas 1985a).<br />

Race cretaceus occupies calcareous grasslands in southern<br />

England; it is large with relatively pale males. Race argus is<br />

found elsewhere.<br />

Distribution: Europe and temperate Asia. In Britain, P. argus<br />

survives in Wales and southern England but is extinct in<br />

Scotland.<br />

97<br />

Population Size: Race cretaceus has suffered a reduction in<br />

range and is only recorded now at Portland Bill (Dorset), where<br />

the remaining colonies are in good health (Warren 1986;<br />

Thomas and Lewington 1991). Race caernensis is thriving,<br />

despite its restricted range in north Wales. In 1983, the peak<br />

emergence was about 250,000 in 10 colonies on the Great<br />

Orme, and about 30,000 in 16 colonies in the Dulas valley: the<br />

total emergence was perhaps three times greater (CD. Thomas<br />

1985b). Numbers were similar at both localities in 1990. An<br />

introduced population was established near Prestatyn in 1983,<br />

and was vigorous by 1990 (Thomas and Harrison 1992).<br />

Heathland populations in north Wales are mostly smaller than<br />

those on the limestone, but one large population (about 40,000<br />

at peak in 1983) occurs on the RSPB Reserve <strong>of</strong> South Stack<br />

Range. P. argus did not decline on heathlands in north Wales<br />

between 1983 and 1990 (Thomas and Harrison 1992). The<br />

mossland population in north Wales contained about 5000<br />

adults at peak in 1983, and was similar in 1990.<br />

Race argus populations occur predominantly on heathland,<br />

where it is mostly a story <strong>of</strong> continuing attrition. Only one<br />

population is left in the Midlands, and just a handful survive in<br />

East Anglia and south Wales. On the Sandlings <strong>of</strong> Suffolk, six<br />

<strong>of</strong> nine colonies contained more than 500 individuals at peak in<br />

1986, yet only one or possibly two still contained this number<br />

by 1990, and two colonies were on the verge <strong>of</strong> extinction: the<br />

total 1990 Sandlings population was estimated to be just 23%<br />

<strong>of</strong> that in 1986, flying over 53% <strong>of</strong> the 1986 area (Ravenscr<strong>of</strong>t<br />

1986,1987,1990, pers. comm.). Similarly, in Devon P. argus<br />

is practically confined to one area <strong>of</strong> heathland (Read 1985). It<br />

is only in the Poole Basin, New Forest, and on the west Surrey<br />

heaths that race argus remains numerous (Thomas and<br />

Lewington 1991). A few populations survive on sand dunes in<br />

Cornwall.<br />

Habitat and Ecology: P. argus occurs on heathlands, calcareous<br />

grassland, sand dunes and mossland, and the larvae feed on a<br />

wide variety <strong>of</strong> hosts in the Leguminosae, Ericaceae and<br />

Cistaceae (CD. Thomas 1985a,b; Jordano et al. in press).<br />

Despite the apparent breadth <strong>of</strong> biotopes and host plants, P.<br />

argus is local because it has other specialised requirements. In<br />

the north, particularly, P. argus is restricted to warm


microclimates: it occurs on sites that contain plenty <strong>of</strong> hot, bare<br />

ground, mostly on south-facing slopes. The eggs are laid at the<br />

margins <strong>of</strong> bare ground and vegetation, and the larvae feed on<br />

the tender growth <strong>of</strong> their host plants. Further south, vegetation<br />

edges seem to be less important, but relatively short successional<br />

vegetation is still favoured (Ravenscr<strong>of</strong>t 1990; Thomas and<br />

Lewington 1991; Jordano et al. in press, N. Armour-Chelu,<br />

pers. cotnm.).<br />

Eggs are laid in midsummer, in response to ants (N. Armour-<br />

Chelu, pers. comm.), and these overwinter and hatch in spring.<br />

The hatchling larvae are attractive to, and are usually picked-up<br />

by, the workers <strong>of</strong> Lasius niger or Lasius alienus, and they are<br />

taken back to the nest (Ravenscr<strong>of</strong>t 1990; Jordano and CD.<br />

Thomas in press). Quite what happens to the first instar larvae<br />

in the nest is unknown, but the later instars feed on foliage above<br />

ground, constantly tended by Lasius (CD. Thomas 1985a;<br />

Ravenscr<strong>of</strong>t 1990; Jordano and CD. Thomas in press). The ants<br />

are pugnacious in defence <strong>of</strong> larvae and, if provoked, will pick<br />

up and retreat with any that are small enough to carry. Pupae can<br />

be found under stones, where they are always tended by Lasius,<br />

and as <strong>of</strong>ten as not they are inside Lasius nests. The emerging<br />

adults are tended by Lasius.<br />

In different populations, P. argus eggs are significantly<br />

associated with different species <strong>of</strong> plants or combinations <strong>of</strong><br />

plants and microhabitats (CD. Thomas 1985a; Read 1985;<br />

Ravenscr<strong>of</strong>t 1990; J. A. Thomas 1991; Jordano et al. in press; N.<br />

Armour-Chelu, pers. comm.): indeed, there are almost as many<br />

different plant associations as populations studied. In contrast,<br />

four studies all show an association <strong>of</strong> P. argus with Lasius ants<br />

(e.g. Ravenscr<strong>of</strong>t 1990; N. Armour-Chelu, pers. comm.).<br />

Suitable conditions occur in Britain where there is a coincidence<br />

<strong>of</strong> young host plant foliage, warm conditions, and adequate<br />

densities <strong>of</strong> Lasius.<br />

Threats: P. argus is threatened by biotope loss and<br />

fragmentation, caused by modern agriculture, afforestation,<br />

urbanisation, etc. For example, in one <strong>of</strong> the best remaining<br />

regions for heathland, the Poole Basin in Dorset, only 14.6% <strong>of</strong><br />

the original heathland survived to 1978: 62% <strong>of</strong> the remaining<br />

fragments are <strong>of</strong>


nearly 50 years in a series <strong>of</strong> limestone habitat patches in the<br />

Dulas valley in north Wales.<br />

Acknowledgements<br />

I thank N. Armour-Chelu, D. Jordano and N. Ravenscr<strong>of</strong>t for<br />

access to unpublished material.<br />

References<br />

DE WORMS, C.M.G. 1949. An account <strong>of</strong> some forms <strong>of</strong> Plebejus argus L.<br />

Rpt. Raven Ent. Nat. Hist. Soc. 1949: 28–30.<br />

HEATH, J., POLLARD, E. and THOMAS, J.A. 1984. Atlas <strong>of</strong> <strong>Butterflies</strong> in<br />

Britain and Ireland. Viking, Harmondsworth.<br />

JORDANO, D., RODRIGUEZ, J., THOMAS, CD. and FERNANDEZ<br />

HAEGER, J. (in press). The distribution and density <strong>of</strong> a lycaenid<br />

butterfly in relation to Lasius ants. Oecologia.<br />

JORDANO, D. and THOMAS, CD. (in press). Specificity <strong>of</strong> an ant-lycaenid<br />

interaction. Oecologia.<br />

RAVENSCROFT, N.O.M. 1986. An investigation into the distribution and<br />

ecology <strong>of</strong> the silver-studded blue butterfly (Plebejus argus L.) in Suffolk:<br />

an interim report. Suffolk Trust for Nature <strong>Conservation</strong>, Ipswich.<br />

RAVENSCROFT, N.O.M. 1987. Plebejus argus colony population estimates<br />

and changes in status in Suffolk 1985–1986. Suffolk Trust for Nature<br />

99<br />

<strong>Conservation</strong>, Ipswich.<br />

RAVENSCROFT, N.O.M. 1990. The ecology and conservation <strong>of</strong> the silverstudded<br />

blue butterfly Plebejus argus L. on the Sandlings <strong>of</strong> East Anglia,<br />

England. Biol. Conserv. 53: 21–36.<br />

READ, M. 1985. The silver-studded blue conservation report. Msc thesis,<br />

Imperial College.<br />

THOMAS, CD. 1985a. Specialisations and polyphagy <strong>of</strong> Plebejus argus<br />

(Lepidoptera: <strong>Lycaenidae</strong>) in North Wales. Ecol. Ent. 10: 325–340.<br />

THOMAS, CD. 1985b. The status and conservation <strong>of</strong> the butterfly Plebejus<br />

argus L. (Lepidoptera: <strong>Lycaenidae</strong>) in North West Britain. Biol. Conserv.<br />

33: 29–51.<br />

THOMAS, CD. 1991. Spatial and temporal variability in abutterfly population.<br />

Oecologia 87: 577–580.<br />

THOMAS, CD. and HARRISON, S. 1992. Spatial dynamics <strong>of</strong> a patchily<br />

distributed butterfly species. J. Anim. Ecol. 61:<br />

THOMAS, J.A. 1991. Rare species conservation: case studies <strong>of</strong> European<br />

butterflies. In: I.F. Spellerberg, F.B. Goldsmith and M.G. Morris (Eds)<br />

The scientific management <strong>of</strong> temperate communities for conservation.<br />

BESSymp. 31: 149–197.<br />

THOMAS, J. and LEWINGTON, R. 1991. The <strong>Butterflies</strong> <strong>of</strong> Britain and<br />

Ireland. Dorling Kindersley, London.<br />

WARREN, M.S. 1986. The status <strong>of</strong> the cretaceus race <strong>of</strong> the silver-studded<br />

blue butterfly, Plebejus argus L., on the Isle <strong>of</strong> Portland. Proc. Dorset<br />

Nat. Hist. Arch. Soc. 108: 153–155.<br />

WEBB, N.R. and HASKINS, L.E. 1980. A ecological survey <strong>of</strong> heathlands in<br />

the Poole Basin, Dorset, England, in 1978. Biol. Conserv. 17: 281–296.


The Zephyr Blue, Plebejus pylaon (Fischer-Waldheim)<br />

M.L. MUNGUIRA and J. MARTIN<br />

Departamento de Biologia (Zoologia), Facultad de Ciencias, Universidad Autonoma de Madrid, Madrid, Spain<br />

Area: Spain, Switzerland, Italy, Hungary, Bulgaria, Romania,<br />

Yugoslavia, Albania, Greece, Turkey, Russia and Asia Minor.<br />

Status and <strong>Conservation</strong> Interest: Status – Hungary:<br />

endangered; Spain, Switzerland, Italy: vulnerable; other<br />

countries: indeterminate, (see Heath 1981; Munguira et al.<br />

1991).<br />

The species, known also as the 'nina del astragalo', is<br />

present over a wide area, but always local and isolated in several<br />

populations some <strong>of</strong> which form different subspecies (species<br />

for some authors, Kudrna 1986, Bálint and Kertész 1990).<br />

Nevertheless, young stages <strong>of</strong> all these different forms are<br />

dependent on Astragalus plants and live on dry steppe-like<br />

habitats. This makes the butterfly rare throughout its range. The<br />

species' biotopes are at present being altered as a result <strong>of</strong> the<br />

change from traditional land uses to more aggressive agricultural<br />

practices, or are being abandoned. The conservation <strong>of</strong> the<br />

species cannot be achieved by simply protecting the sites, due<br />

to the seral character <strong>of</strong> P. pylaon habitats, and would need<br />

certain management practices in order to maintain desirable<br />

ecological features.<br />

Taxonomy and Description: The taxonomic level <strong>of</strong> the forms<br />

under consideration is not clear at the moment and needs further<br />

study. Plebejus vogelii Oberthür and particularly P. martini<br />

Allard, are true species for most <strong>of</strong> the authors due to their<br />

morphological and ecological features (Higgins and Hargreaves<br />

1983, Bálint and Kertész 1990). Nevertheless, the taxa included<br />

under the name pylaon are treated by some authors as distinct<br />

species. Kudrna (1986) for example splits the group into four<br />

different species in Europe: hesperica Rambur,pylaon, sephirus<br />

and trappi Verity. Bálint and Kertész (op. cit.) group the<br />

European forms in these same four taxa, but they split each<br />

group into several biogeographical sub-groups without naming<br />

them as separate species. Within these so-called species a large<br />

number <strong>of</strong> subspecies have been described, but at least with the<br />

Spanish ones (group hesperica above) we consider all the<br />

subspecies as synonyms (Munguira 1989), and this certainly<br />

may be the case in other groups.<br />

100<br />

Distribution: The species is present from central Spain to the<br />

surroundings <strong>of</strong> the Baikal Lake. In Spain it lives in the central<br />

Plains, Iberian Mountains and near Sierra Nevada in Granada<br />

Province in a total <strong>of</strong> 37 UTM squares <strong>of</strong> 10 x 10km (Munguira<br />

1989, Munguira and Martin 1989). In Italy and Switzerland<br />

subspecies trappi is also very local. In the last country it is only<br />

found in 16 UTM squares (Gonseth 1987) in the Valais. It is also<br />

very local in Hungary (Bálint and Kertész 1990) and only<br />

present in 22 UTM squares in Yugoslavia (Jaksic 1988).<br />

Population Size: In Spain the populations are local but the<br />

number <strong>of</strong> adults is high. Many plants can have up to 20 eggs,<br />

giving adult yields in each population <strong>of</strong> several thousand<br />

individuals. After overwintering, larval numbers may reach the<br />

level <strong>of</strong> 2–3 per plant. Due to the local character <strong>of</strong> the<br />

populations, some localities are sensitive to extinction events.<br />

In Sierra de Alfcar, the type locality <strong>of</strong> Spanish hesperica, the<br />

butterfly has never been collected since its description in 1839.<br />

In central Spain a colony was partially destroyed by a limestone<br />

Plebejus pylaon, male, Camporeal (photo by M. Munguira).


pit (Gomez-Bustillo 1981) but at present the species is still on<br />

the site although in reduced numbers. Other localities still<br />

support important numbers <strong>of</strong> the species but are vulnerable to<br />

human impacts.<br />

Habitat and Ecology: The species always occurs on disturbed<br />

Quercus rotundifolia forests (encinares) in the Iberian Peninsula.<br />

Serai communities are maintained on dirt road verges, quarries,<br />

or on land disturbed by overgrazing. The geological substrate<br />

is clay or limestone. The soil is exposed in approximately 75%<br />

<strong>of</strong> the surface <strong>of</strong> the places where the foodplant (Astragalus<br />

alopecuroides) grows. The vegetation is formed by strong<br />

rooted perennials and shrubs specialised to live in habitats with<br />

very poor soil conditions. Altitude ranges from 400 to 1400m,<br />

and rainfall is always scarce (between 400 and 500mm per year)<br />

in every locality studied.<br />

Eggs are laid on the leaflets <strong>of</strong> the foodplant in May or early<br />

June. After a week the first instar larvae begin to feed on the<br />

parenchyma <strong>of</strong> the leaves, leaving characteristic eye-like damage<br />

on the plant. At the beginning <strong>of</strong> July the larvae undergo the<br />

second moult and build a silken refuge on the base <strong>of</strong> the<br />

foodplant in which they spend all the rest <strong>of</strong> the summer and the<br />

winter (aestivating and overwintering). In the following March<br />

the larvae begin to feed on the young shoots <strong>of</strong> the plant and<br />

develop quickly until reaching the full-grown condition (fifth<br />

Habitat <strong>of</strong> P. pylaon, Camporeal, May 1991 (photo by M. Munguira).<br />

101<br />

instar) in March–April. They pupate in the ground and emerge<br />

as adults 20 to 30 days later.<br />

The last two larval instars are invariably tended by ants<br />

belonging to several species at each locality. We have found<br />

attending ants <strong>of</strong> eight different species <strong>of</strong> the genera Formica,<br />

Camponotus, Crematogaster and Plagiolepis. Larvae have<br />

dorsal nectary organs (Newcomer's gland) and tentacle organs.<br />

Tentacles are displayed when the larva is in danger, and the<br />

attending ants become excited and attack any potential enemy<br />

when this happens. This behaviour probably defends larvae<br />

against parasitoids and in fact, despite having reared almost 40<br />

larvae to adults, we never obtained parasitoids from the species.<br />

Relationships with ants are facultative, and the whole life cycle<br />

can be completed without the ants' presence.<br />

The butterfly is present in all the localities where we have<br />

seen its foodplant. Nevertheless, the plant also exists in southern<br />

France where pylaon has never been recorded. In the high<br />

altitude localities <strong>of</strong> the Iberian Mountains the foodplant is the<br />

Iberian-African A. turolensis (Sheldon 1913). The biology and<br />

ecology <strong>of</strong> subspecies trappi and sephirus are similar (SBN<br />

1987; Balint and Kertész 1990). The foodplant <strong>of</strong> trappi is<br />

Astragalus exscapus, and the butterfly appears a month later<br />

than in Spain, making the whole life cycle a month late. The<br />

habitat in the Valais consists <strong>of</strong> dry grasslands on rocky limestone<br />

slopes.


Threats: Most lowland populations are endangered by<br />

urbanisation, because they live on flat areas suitable as building<br />

sites. In the Tagus Valley the species lives in surrounding small<br />

olive groves, and modern agriculture practices can also destroy<br />

this habitat, reducing headlands and hedges to a minimum as a<br />

consequence <strong>of</strong> management intensification. While small fields<br />

maximise the border area, increases in the grove area are<br />

reducing the headlands and hedges.<br />

Some <strong>of</strong> the highland populations are threatened by road<br />

developments (SBN 1987), or because they are potential areas<br />

for pine plantations in the Spanish Iberian Mountains (Munguira<br />

1989).<br />

Most <strong>of</strong> the areas in which the butterfly lives have seral<br />

vegetation communities. Rabbit or sheep grazing is probably<br />

necessary to maintain the proper habitat in Spain (Munguira<br />

1989). On the other hand in Switzerland, sheep overgrazing<br />

probably endangers the foodplant suitability (SBN 1987).<br />

<strong>Conservation</strong>: The species is present at the moment in several<br />

Natural and National Parks: Tshatkal Reserve (Uzbekistan),<br />

Galicica National Park (southern Yugoslavia), Gran Paradiso<br />

National Park (Val d'Aosta, Italy) and El Regajal Reserve<br />

(Madrid, Spain). This by no means assures proper protection<br />

for the butterflies, because National Parks and Reserves are<br />

normally created to protect the larger mammals and birds. At<br />

least with these reserves, the four European subspecies are<br />

protected in theory. The real conservation <strong>of</strong> the species can<br />

only be achieved when proper management is maintained on at<br />

least several suitable hectares on each reserve. This management<br />

involves continuing with traditional land uses such as extensive<br />

grazing. Nevertheless, the exact grazing regime and the required<br />

sheep and goat stocking rates are not known properly, and<br />

require further study.<br />

The Regajal Reserve, one <strong>of</strong> the Spanish localities where<br />

pylaon is present, illustrates the conflicts between insect<br />

conservation and development. The area was suggested as a<br />

Lepidoptera Reserve in the mid-seventies (Viedma and Gomez-<br />

Bustillo 1976). Later on a project to build a roundabout crossing<br />

the area was conceived, receiving the criticism <strong>of</strong> amateur and<br />

pr<strong>of</strong>essional entomologists alike. The site is not only valuable<br />

102<br />

for its rarities, but also because it has always been a collecting<br />

site for most <strong>of</strong> Spain's leading entomologists. The motorway<br />

was finally built in 1989, crossing the reserve through its very<br />

centre. The impact <strong>of</strong> the motorway is certainly serious: for<br />

example the rivers have changed their regime, causing several<br />

serious floods in 1990. The damage to insect populations will<br />

never be known because a suitable impact assessment was not<br />

done before building the motorway. After all these events, the<br />

area is receiving a protected status by regional authorities, and<br />

will probably be the first Spanish Reserve created mainly to<br />

protect Lepidoptera.<br />

References<br />

BÁLINT, Z. and KERTÉSZ, A. 1990. A survey <strong>of</strong> the subgenus Plebejides<br />

(Sauter, 1968) – preliminary revision. Linn. Belg. 12: 190–224.<br />

GOMEZ-BUSTILLO, M.R. 1981. Protection <strong>of</strong> Lepidoptera in Spain. Beih.<br />

Ver<strong>of</strong>f. Natursch. Landschaft. Bad.-Wurt. 21: 67–72.<br />

GONSETH, Y. 1987. Atlas de distribution des papillons diurnes de Suisse<br />

(Lepidoptera Rhopalocera). CSCF, Neuchatel.<br />

HEATH, J. 1981. Threatened Rhopalocera (butterflies) in Europe. Council <strong>of</strong><br />

Europe, Strasbourg.<br />

HIGGINS, L.G. and HARGREAVES, B. 1983. The <strong>Butterflies</strong> <strong>of</strong> Britain and<br />

Europe. Collins, London.<br />

JAKSIC, P. 1988. Provisional distribution maps <strong>of</strong> the butterflies <strong>of</strong> Yugoslavia<br />

(Lepidoptera, Rhopalocera). Societas Entomologica Jugoslavia, Zagreb.<br />

KUDRNA, 0.1986. <strong>Butterflies</strong> <strong>of</strong> Europe. Vol. 8. Aspects <strong>of</strong> the <strong>Conservation</strong><br />

<strong>of</strong> <strong>Butterflies</strong> in Europe. Aula-Verlag, Wiesbaden.<br />

MUNGUIRA, M.L. 1989. Biologia y biogeografia de los licenidos ibericos en<br />

peligro de extincion (Lepidoptea, <strong>Lycaenidae</strong>). Ediciones Universidad<br />

Autonoma de Madrid, Madrid, Spain.<br />

MUNGUIRA, M.L. and MARTIN, J. 1989. <strong>Biology</strong> and conservation <strong>of</strong> the<br />

endangered lycaenid species <strong>of</strong> Sierra Nevada, Spain. Nota lepid. 12<br />

(Suppl. 1): 16–18.<br />

MUNGUIRA, M.L., MARTIN, J. and REY, J.M. 1991. Use <strong>of</strong> UTM maps to<br />

detect endangered lycaenid species in the Iberian Peninsula. Nota lepid.<br />

Suppl. 2: 45–55.<br />

SBN. 1987. Tagfalter und ihre Lebensraume. Schweizerischer Bund fur<br />

Naturschutz, Basel.<br />

SHELDON.W.G. 1913. Lepidoptera at Albarracin in May and June, 1913.<br />

Entomologist 46: 283–9, 309–13, 328–32.<br />

VIEDMA, M.G. and GOMEZ-BUSTILLO, M.R. 1976. Libro Rojo de los<br />

lepidopteros ibericos. ICONA, Madrid.


The Pannonian Zephyr Blue, Plebejus sephirus kovacsi Szabó<br />

Z. BÁLINT<br />

Zoological Department, Hungarian Natural History Museum, Baross utca 13, Budapest, H–1008, Hungary<br />

Country: Hungary.<br />

Status and <strong>Conservation</strong> Interest: Status – rare.<br />

The butterfly was found in 1949, in the close vicinity <strong>of</strong><br />

Budapest and it was thought that this population was the only<br />

Hungarian one. Several new colonies were discovered in the<br />

1980s, but all were within a 30km radius <strong>of</strong> the first one. These<br />

colonies represent the most western as well as the most northern<br />

occurrences <strong>of</strong> the species. Very recently a new population has<br />

been found in northeast Hungary showing continuity towards<br />

the Transylvanian (Romania) colonies.<br />

Taxonomy and Description: P. sephirus is a member <strong>of</strong> a<br />

western palaearctic group <strong>of</strong> lycaenids <strong>of</strong> central Asian<br />

xeromontane origin (Báiint and Kertész 1990a; Bálint 1991a).<br />

Several described subspecies <strong>of</strong> sephirus exist in the Carpathian<br />

Basin (central Hungary: ssp. kovacsi Szabó, 1954 (= foticus<br />

Szabó, 1956); Banat (Voivodina): ssp. uhryki Rebel, 1911;<br />

Transylvania: ssp. proximus Szabó, 1954), some which are<br />

most probably synonyms <strong>of</strong> the nominate subspecies sephirus<br />

Frivaldszky, 1835 occurring in the Balkans.<br />

Distribution: Very restricted in two isolated parts <strong>of</strong> the<br />

country. Eight small colonies can be found north <strong>of</strong> Budapest,<br />

one in Tokaj, northeast Hungary, found very recently (1991) by<br />

Pr<strong>of</strong>essor Varga.<br />

Population Size: All colonies are strongly isolated, and the<br />

structure <strong>of</strong> the populations is closed. There is no possibility <strong>of</strong><br />

interchange <strong>of</strong> adults even between the closest populations,<br />

whilst they are separated by heavily cultivated agricultural<br />

regions or human settlements. One small population which is<br />

divided by a double rail track was studied by the capturerecapture<br />

method. The results show that the adults are strongly<br />

stenochorous, with the life history <strong>of</strong> the species closely<br />

associated with the larval foodplant.<br />

Two <strong>of</strong> the larger populations have been monitored during<br />

the last two years. All known colonies in the vicinity <strong>of</strong><br />

Budapest have been estimated by counts <strong>of</strong> adults during the<br />

last three years. The major two colonies contained about 700–900<br />

103<br />

butterflies. Three smaller colonies had fewer than 100<br />

individuals.<br />

Habitat and Ecology: Colonies are strongly confined to the<br />

association Astragalo-Festucetum rupicolae, a typical habitat<br />

<strong>of</strong> forest steppe on loose calciferous soil. There is one generation<br />

each year with a very short flight period <strong>of</strong> adults from the<br />

middle <strong>of</strong> May to the beginning <strong>of</strong> June. Adults are active only<br />

when the air temperature is above 25°C, and the wind is not<br />

strong. Eggs are laid singly on the larval foodplant, Astragalus<br />

exscapus L. Caterpillars hatch after about ten days. After a short<br />

feeding period they retreat to ant chambers at the base <strong>of</strong> the<br />

plants. The diapause lasts from about the end <strong>of</strong> June (because<br />

<strong>of</strong> the long hot summer) to the middle <strong>of</strong> the following spring,<br />

about the middle <strong>of</strong> April. All older instars are tended by ants,<br />

so Plebejus sephirus is a steadily myrmecophilous species. The<br />

following ant species are involved: Tetramorium (caespitum<br />

gr.); Formica pratensis; Camponotus aethiops; and Lasius<br />

(alienus gr.) (Fiedler 1991). Caterpillars pupate in the upper<br />

end <strong>of</strong> the ant chambers. The pupal stage lasts about ten days<br />

and the pupae are also tended by ants. Adults <strong>of</strong> P. sephirus take<br />

nectar mainly from Dianthus giganteiformis Borb. ssp.<br />

pontederae Kern., endemic to the Pannonian region.<br />

Threats: The present scattered Hungarian distribution <strong>of</strong> the<br />

species shows that current distribution is mainly the result <strong>of</strong><br />

human activity: the loose soil <strong>of</strong> the central Carpathian Basin<br />

has been cultivated from early historical times. The major<br />

recent threats have been: (1) the establishment <strong>of</strong> new housing<br />

estates; (2) illegal rubbish heaps; (3) motor-cross activities; (4)<br />

afforestation; (5) natural succession; (6) overcollecting.<br />

<strong>Conservation</strong>: The first discovered habitat <strong>of</strong> P. sephirus,<br />

namely Somlyóhegy near Fót, has been a nature reserve since<br />

1953. Two larger areas, where the strongest colonies exist, will<br />

be protected in the very near future. The butterfly and its larval<br />

hostplant are protected by Hungarian law (Bálint and Kertész<br />

1990b). A paper on the conservation and management <strong>of</strong> P.<br />

sephirus was ordered by the Hungarian authorities and was


drawn up (Bálint 1991b), but there is no financial support to<br />

undertake the practical measures suggested to ensure good<br />

management. The recommendations for the P. sephirus<br />

populations in the Budapest area are as follows:<br />

1. Protection <strong>of</strong> all colonies during the flight period <strong>of</strong> the<br />

species;<br />

2. Monitoring <strong>of</strong> all populations;<br />

3. Two important habitats with strong P. sephirus colonies<br />

could be attached to the Somlyohegy Nature Reserve, which<br />

could be a special reserve for P. sephirus, where it is<br />

necessary to arrest natural succession and negative human<br />

influences;<br />

4. Stronger publicity for conservationist activities.<br />

104<br />

References<br />

BÁLINT, Zs. 1991a. (A xeromontane lycaenid butterfly: Plebejus pylaon<br />

(F.W., 1832) and its relatives), I. Jan. Pann. Muz. Évk. 35: 33–69 (in<br />

Hungarian).<br />

BÁUNT, ZS. 1991b. (The ecology and conservation <strong>of</strong> Plebejus sephirus<br />

(Frivaldszky, 1835) in Hungary). Budapest (in Hungarian).<br />

BALINT, Zs. and KERTÉSZ, A. 1990a. A survey <strong>of</strong> the subgenus Plebejides<br />

Sauter, 1968 – preliminary revision. Linneana Belgica 12: 190–224.<br />

BÁLINT, Zs. and KERTÉSZ, A. 1990b. The conservation <strong>of</strong> Plebejus sephirus<br />

(Frivaldszky, 1835) in Hungary. Linneana Belgica 12: 254–273.<br />

FIEDLER, K. 1991. Systematic, evolutionary, and ecological implications <strong>of</strong><br />

myrmecophily within the <strong>Lycaenidae</strong> (Insecta: Lepidoptera:<br />

Papilionoidea). Bonn. Zool. Monog. 31, 210 pp.


The threatened lycaenids <strong>of</strong> the Carpathian Basin, east-central<br />

Europe<br />

Zsolt BÁLINT<br />

Hungarian Natural History Museum, Zoological Department, Budapest, Baross utca 13. H–1008<br />

The following species are notable <strong>Lycaenidae</strong> which occur in<br />

the region. Each is treated as a discrete account and these 17<br />

taxa collectively indicate the main lycaenid conservation needs<br />

in the Carpathian Basin, and the eastern adjacent region <strong>of</strong><br />

Romania. The references are given for the whole account rather<br />

than each species separately.<br />

Aricia macedonica isskeutzi Balogh<br />

Country: Hungary.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

A subspecies confined to a very small region <strong>of</strong> northern<br />

Hungary.<br />

Taxonomy and Description: Close to Aricia allous (Geyer)<br />

(see Geiger 1988) but differing in some minor genitalic and<br />

superficial characters (Varga 1968).<br />

Distribution: Aricia macedonica Verity distributed in the<br />

Balkans is an allopatric group <strong>of</strong> taxa, distinct from that <strong>of</strong> A.<br />

artaxerxes (Fabricius) (Britain), A. allous (Alps and central<br />

Europe) and A. inhonora (Jachontov) (Scandinavia, Russia).<br />

The subspecies issekutzi is the most northern and western<br />

representative <strong>of</strong> the macedonica-complcx, and it can be found<br />

in the Karst <strong>of</strong> Torna (northern Hungary and southern Slovakia)<br />

and in the Bükk Mountains (northern Hungary).<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Xerothermophilous, univoltine (from<br />

late June to early August). Stenotopic, inhabiting clearings <strong>of</strong><br />

forest steppes, rocky and dry grasslands. Caterpillar steadily<br />

my rmecophilous (Fiedler 1991). Larval foodplants Helianthemum<br />

ovatum (Viv.) Dunal and Geranium sanguineum L.<br />

Threats: The populations are strongly isolated and threatened<br />

by intensification <strong>of</strong> grassland management, afforestation,<br />

urbanisation and tourism.<br />

105<br />

<strong>Conservation</strong>: Most <strong>of</strong> the Hungarian habitats can be found in<br />

the territories <strong>of</strong> Aggtelek National Park and Bükk National<br />

Park. Thorough ecological studies are necessary to continue the<br />

work <strong>of</strong> Varga (1968).<br />

Aricia eumedon (Esper)<br />

Country: Hungary.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

A protected species, which is only known from northern<br />

Hungary. The example recorded from Transdanubia (W.<br />

Hungary), collected at the beginning <strong>of</strong> the century, has never<br />

been confirmed.<br />

Taxonomy and Description: All the central European<br />

populations are identical (see Geiger 1988).<br />

Distribution: Transpalearctic. The species is widely distributed<br />

in the brook or river valleys <strong>of</strong> the Carpathians (Slovakia), but<br />

it is absent from the central part <strong>of</strong> this range.<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Hygrophilous, univoltine (mainly July).<br />

Stenotopic, the colonies can be found in wetlands along water<br />

courses with luxuriant vegetation <strong>of</strong> the larval foodplant<br />

Geranium palustre (L.). Caterpillars are steadily<br />

myrmecophilous (Fiedler 1991). The main nectar source <strong>of</strong> the<br />

adults is also the above-mentioned purple flowered Geranium<br />

species.<br />

Threats: Natural succession, overcollecting.<br />

<strong>Conservation</strong>: All the Hungarian colonies, which constitute a<br />

single metapopulation, are in a small mountain stream valley<br />

situated in the Aggtelek Natural Park. Patrolling <strong>of</strong> the known<br />

sites is needed along with a study <strong>of</strong> the ecosystem <strong>of</strong> the stream<br />

valley.


Cupido osiris (Meigen)<br />

Country: Hungary.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered.<br />

A protected species. Only three Hungarian sites are known,<br />

and these are the most northern permanent populations <strong>of</strong> this<br />

species.<br />

Taxonomy and Description: The Hungarian populations do<br />

not differ from the central European ones (see Geiger 1988).<br />

Distribution: Mediterranean. Very rare in Slovakia, several<br />

populations in Transylvanian Basin, Romania (Bálint 1985a).<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Xerothermophilous, stenotopic, probably<br />

bivoltine. The Hungarian habitats include abandoned orchards<br />

and vineyards. Caterpillars are steadily myrmecophilous (Fiedler<br />

1991). Larval foodplants are Colutea arborescens (L.) and<br />

Onobrychis viciaefolia Scop..<br />

Threats: Only one <strong>of</strong> the three populations has been monitored<br />

in the last years, most probably the most endangered one.<br />

Factors giving concern include: earthworks as a result <strong>of</strong> openpit<br />

mining; air pollution from a lime factory; and the recultivation<br />

<strong>of</strong> abandoned orchards and vineyards.<br />

<strong>Conservation</strong>: The conservation <strong>of</strong> the species in Hungary has<br />

not yet been achieved. Patrolling and study <strong>of</strong> the known<br />

populations, together with the monitoring <strong>of</strong> suitable habitats<br />

for new sites are urgently required.<br />

Jolana iolas (Ochsenheimer)<br />

Country: Hungary.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

A protected species, J. iolas was discovered in Hungary.<br />

The habitat <strong>of</strong> the original specimens was recently destroyed by<br />

urbanisation. Many known colonies have disappeared in recent<br />

decades.<br />

Taxonomy and Description: The Hungarian populations do<br />

not differ from the Central European ones (Geiger 1988).<br />

Distribution: Mediterranean. The northern limit <strong>of</strong> the species<br />

range can be found in Hungary. The southern Slovakian records<br />

have not been confirmed recently (Kulfan and Kulfan 1991).<br />

Population Size and Status: Not known.<br />

106<br />

Habitat and Ecology: Xerothermophilous, univoltine<br />

(May-July), stenotopic. The known habitats are mainly<br />

abandoned vineyards or Pannonian karst oak-scrubs. Caterpillars<br />

are presumed to be moderately myrmecophilous (Fiedler 1991)<br />

and the larval foodplant is Colutea arborescens (L.) (Uhrik-<br />

Mészáros, 1948; Szabó, 1956). Adults fly around the bushes <strong>of</strong><br />

the larval foodplant, which is also their main nectar source.<br />

Threats: Afforestation (planting <strong>of</strong> Pinus nigra Arn.),<br />

urbanisation (a lot <strong>of</strong> habitats have been built upon) and<br />

overcollecting.<br />

<strong>Conservation</strong>: Some colonies can be found in Landscape<br />

Protection Areas. Their habitats need full protection. The<br />

ecology <strong>of</strong> the species must be studied.<br />

Maculinea alcon (Denis & Schiffermuller)<br />

Country: Hungary.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered.<br />

The Hungarian populations have almost totally disappeared<br />

and only a very few colonies remain in the western part <strong>of</strong> the<br />

country.<br />

Taxonomy and Description: The Hungarian populations are<br />

identical with the European ones (see Geiger 1988).<br />

Distribution: West Palaearctic. The most easterly populations<br />

can be found in Hungary and in Romania (Transylvania).<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Hygrophilous, univoltine (late<br />

July-beginning <strong>of</strong> August), stenotopic. Larval hostplant is<br />

Gentiana pneumonanthe (L.). Caterpillars are obligately<br />

myrmecophilous (Fiedler 1991). Other details concerning the<br />

Hungarian populations are not known.<br />

Threats: Wetland drainage.<br />

<strong>Conservation</strong>: All the remaining populations must be<br />

discovered and studied. Some already known colonies can be<br />

found in Landscape Protection Areas. The whole ecosystem<br />

(incorporating the effects <strong>of</strong> traditional agricultural practices)<br />

<strong>of</strong> the habitats must be protected.


Maculinea nausithous (Bergstrasser)<br />

Country: Hungary.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered.<br />

A protected species. Serious loss <strong>of</strong> habitat through wetland<br />

drainage and intensification <strong>of</strong> grassland management has<br />

taken place.<br />

Taxonomy and Description: The Hungarian populations are<br />

identical with the European ones (see Geiger 1988).<br />

Distribution: West Palaearctic. The most easterly populations<br />

can be found in the western part <strong>of</strong> the country (Transdanubia).<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Hygrophilous, univoltine (late<br />

July-beginning <strong>of</strong> August), presumed stenotopic. Larval<br />

hostplant is Sanguisorba <strong>of</strong>ficinalis(L.). Caterpillars are<br />

obligately myrmecophilous (Fiedler 1991). Other details<br />

concerning the Hungarian populations are not known.<br />

Threats: Much <strong>of</strong> the suitable habitat for this species has<br />

become the victim <strong>of</strong> wetland drainage and intensification <strong>of</strong><br />

grassland management.<br />

<strong>Conservation</strong>: All the remaining populations must be<br />

documented and measured and the ecology <strong>of</strong> the species in<br />

Hungary must be studied. Some populations are in the Landscape<br />

Protection Area and in the Ferto-tó National Park, where the<br />

whole ecosystem must be protected effectively.<br />

Maculinea sevastos limitanea Bálint<br />

Country: Romania.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

Endemic to the eastern Carpathians. The populations are<br />

very scattered.<br />

Taxonomy and Description: Resembles Maculinea alcon, but<br />

the wing shape is more expanded, and the underside darker<br />

brown somewhat similar to M. nausithous (Bálint 1986).<br />

Distribution: An east Mediterranean species, its most western<br />

and northern populations can be found in Transylvania<br />

(Romania) (see Kudrna 1986).<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Xerothermophilous, univoltine (July).<br />

Caterpillars are presumed to be obligately myrmecophilous.<br />

107<br />

The larval hostplant is Gentiana crutiata (L.). The species<br />

occurs in the same habitats as Parnassius apollo transylvanicus<br />

(Schweitzer, 1912), and Polyommatus dory las magna (Bálint<br />

1985). Ecologically very similar to the transpalaearctic M.<br />

xerophila Berger.<br />

Threats: Afforestation, intensification <strong>of</strong> grassland management<br />

and tourism.<br />

<strong>Conservation</strong>: The unique ecosystem <strong>of</strong> Békás-szoros (Cheile<br />

Bicazului, the type locality <strong>of</strong> the subspecies) with its<br />

surroundings in the eastern Carpathians could be a National<br />

Park in Romania, where the traditional grassland and forest<br />

management could be maintained. Tourists could be restricted<br />

to indicated paths rather than be allowed to ramble freely.<br />

Plebejus (Lycaeides) idas (L.)<br />

Country: Hungary.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

The species is very little known in Hungary, and only<br />

scattered faunistic records exist. Its typical habitat (heath covered<br />

with Calluna vulgaris (L.) Hull. and Vaccinium myrtillus L.) is<br />

very rare.<br />

Taxonomy and Description: The Hungarian populations<br />

belong to one <strong>of</strong> the central European subspecies described as<br />

stempfferschmidti Beuret, which is most probably identical<br />

with the German or the other central European populations<br />

(Geiger 1988). The species needs a comprehensive taxonomic<br />

revision.<br />

Distribution: Holarctic. Most probably the Mediterranean<br />

(Balkans, Serbia, Croatia)'idas' populations represent another<br />

species, so the Hungarian colonies seem to be boundary<br />

populations.<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Xerothermophilous, bi- or trivoltine,<br />

presumed stenotopic. Caterpillars are regularly or obligately<br />

myrmecophilous (Fiedler 1991). Its typical habitat is heath<br />

covered with Calluna vulgaris (L.) Hull. and Vaccinium myrtillus<br />

L.). Other details (such as larval foodplants in other localities,<br />

etc.) are unknown from Hungary. Sarothamnus scoparius (L.)<br />

Wimm. is known as the larval foodplant in northeastern Hungary,<br />

on a site with atlantic influences.<br />

Threats: Intensification <strong>of</strong> grassland and forestry management.<br />

<strong>Conservation</strong>: Only one stable and strong population is known.<br />

It is situated on a xerophytic, silicate grass steppe <strong>of</strong> an acid<br />

mountain slope, partly covered with scrub (Quercus cerris L.


and Q. petraea (Mattuschka) Lieblein) in northeast Hungary<br />

(Bálint 1991). This unique habitat was strongly disturbed by the<br />

opening <strong>of</strong> a new forestry service road, the changing <strong>of</strong> the<br />

water-course system and the partial destruction <strong>of</strong> the grassland.<br />

Field research on this species is urgently required in Hungary<br />

to determine if it exists at other sites and to gather basic<br />

information on the species.<br />

Polyommatus (Agrodiaetus) admetus<br />

(Esper)<br />

Country: Hungary.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

The species was described on the basis <strong>of</strong> Hungarian<br />

specimens. A typical forest steppe species, many <strong>of</strong> its habitats<br />

were destroyed in the last decade. Very local in the country and<br />

in the Carpathian Basin.<br />

Taxonomy and Description: The species is a member <strong>of</strong> the<br />

ripartii (Freyer)-group, which consists <strong>of</strong> many very closely<br />

related allopatric species (Geiger 1988). The group was analysed<br />

by De Lesse (1960).<br />

Distribution: An east Mediterranean species. The species has<br />

the western and northern European limits <strong>of</strong> its range in Hungary.<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Xerothermophilous, univoltine (late<br />

June–July), stenotopic. Caterpillar presumed steadily<br />

myrmecophilous (Fiedler 1991). Larval hostplant in Hungary:<br />

Onobrychis arenaria (Kit.) DC. (Szabó 1956). Other details are<br />

not known.<br />

Threats: Major threats are afforestation, recultivation,<br />

urbanisation and air pollution.<br />

Many habitats have been lost through the recultivation <strong>of</strong><br />

abandoned orchards and vineyards or were destroyed when<br />

built over very recently (especially in the area surrounding<br />

Budapest where the species is now extinct). Some colonies<br />

have disappeared because <strong>of</strong> artificially planted Pinus nigra<br />

Arn. (Pest County: Esztergom).<br />

<strong>Conservation</strong>: Only two populations have been confirmed<br />

recently; one <strong>of</strong> them is situated in the territory <strong>of</strong> Aggtelek<br />

National Park. The second population is strongly threatened by<br />

a lime factory (Pest County: Vac). Since the ecology <strong>of</strong> the<br />

species is totally unknown, field studies are urgently required.<br />

Field surveys are also essential since it seems very likely that<br />

more populations exist.<br />

108<br />

Polyommatus (Agrodiaetus) damon (Denis<br />

& Schiffermüller)<br />

Country: Hungary.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered.<br />

The species has been recorded only from two Hungarian<br />

localities (Bálint 1991), and one <strong>of</strong> these has not been confirmed<br />

in recent years. The single remaining population is the only<br />

known confirmed one in the whole Carpathian Basin.<br />

Taxonomy and Description: The Hungarian population is<br />

identical to those <strong>of</strong> central Europe (Geiger 1988, p.388).<br />

Distribution: Transpalearctic (from Iberian peninsula to<br />

Mongolia), but the range consists <strong>of</strong> very scattered populations.<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Xerothermophilous, univoltine (July,<br />

beginning <strong>of</strong> August), stenotopic. Caterpillar nocturnal feeder,<br />

steadily myrmecophilous (Fiedler 1991). Larval hostplant<br />

Onobrychis viciaefolia L. (pers. obs.). Other details are not<br />

known.<br />

Threats: The single remaining population is in the territory <strong>of</strong><br />

Budapest, and this site is the most popular centre for picnics,<br />

weekend activities and winter sports. These recreational activities<br />

represent a serious threat to the species.<br />

<strong>Conservation</strong>: Fencing <strong>of</strong>f the habitat to safeguard the species<br />

is not possible in this area and the only long-term solution for<br />

the species in Hungary seems to be translocation to another<br />

suitable habitat (possibly in a protected area). Suggested field<br />

studies in conjunction with a proposal for the conservation <strong>of</strong><br />

P. damon are under preparation.<br />

Polyommatus (Vacciniina) optilete (L.)<br />

Country: Slovakia.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered.<br />

The only recently known population <strong>of</strong> the species in the<br />

Carpathian Basin can be found in northwest Slovakia.<br />

Taxonomy and Description: The Slovakian populations do<br />

not differ from those <strong>of</strong> central Europe (Geiger 1988).<br />

Distribution: Holarctic. It has been recorded from several<br />

northern Carpathian localities, but only one site has been<br />

confirmed recently (Kulfan and Kulfan 1991).<br />

Population Size and Status: Not known.


Habitat and Ecology: Tyrpophil, univoltine (July), stenotopic.<br />

Caterpillar myrmecoxenous (Fiedler 1991). The larval hostplant<br />

in the Alps is mainly Vaccinium L. species (Geiger 1988). No<br />

details concerning the Slovakian populations are known.<br />

Threats: Natural succession and wetland drainage.<br />

<strong>Conservation</strong>: It would be important to protect the peat bogs<br />

and to study the ecology <strong>of</strong> the Slovakian populations.<br />

Polyommatus eroides (Frivaldszky)<br />

Country: Slovakia.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered.<br />

In the Carpathians, records <strong>of</strong> the species exist only from<br />

Slovakia. The species has not been reported from Hungary, or<br />

even from Romania.<br />

Taxonomy and Description: Resembling Polyommatus eros<br />

(Ochsenheimer) (Geiger 1988), but larger with a deeper blue<br />

upperside ground colour and a somewhat wider black border.<br />

Distribution: An east Mediterranean species, perhaps endemic<br />

for the Balkans. The Slovakian populations were connected<br />

with the Moravian ones, but both <strong>of</strong> them have become strongly<br />

isolated in recent times and are very close to extinction, most<br />

probably by human influences (see Králicek and Povolny<br />

1957).<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Xerothermophilous, univoltine (July),<br />

stenotopic. Caterpillar presumed to be steadily my rmecophilous<br />

(Fiedler 1991). Other aspects <strong>of</strong> the ecology <strong>of</strong> the species are<br />

totally unknown.<br />

Threats: The records are very scattered, which suggests that<br />

the populations are on the way to disappearing. Kulfan and<br />

Kulfan (1991) grouped the species into a complex <strong>of</strong><br />

xerothermophilous butterflies which are threatened by several<br />

harmful factors, including recultivation <strong>of</strong> their habitats<br />

(abandoned orchards and vineyards), burning fields,<br />

overgrazing, building-over, illegal rubbish-dumping,<br />

earthworks, afforestation and natural succession.<br />

<strong>Conservation</strong>: Ecological and taxonomic investigations are<br />

urgently required.<br />

109<br />

Polyommatus dorylas magna (Bálint)<br />

Country: Romania.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

Endemic to the eastern Carpathians.<br />

Taxonomy and Description: Much larger than the nominate<br />

race, with a very wide and conspicuous antemarginal white<br />

border (Bálint 1985b).<br />

Distribution: P. dorylas is distributed in the western Palaearctic<br />

region. This subspecies is known only in the mountain system<br />

<strong>of</strong> the eastern Carpathians (Romania and Carpat-Ukraine,<br />

?Slovakia).<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Xerothermophilous, univoltine (July<br />

and beginning <strong>of</strong> August). Caterpillar steadily myrmecophilous<br />

(Fiedler 1991). Larval hostplant and main nectar source <strong>of</strong> the<br />

imagines: Anthyllis vulneraria L. (pers. obs.). The species lives<br />

in the same habitats as Parnassius apollo transylvanicus<br />

(Schweitzer, 1912) and Maculinea sevastos limitanea (Bálint<br />

1985).<br />

Threats: Afforestation, intensification <strong>of</strong> grassland<br />

management, tourism.<br />

<strong>Conservation</strong>: The same suggestions as those advanced for M.<br />

sevastos limitanea apply equally to this taxon.<br />

Lycaena helle (Denis & Schiffermuller)<br />

Country: Romania.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered.<br />

The species is endangered in the western part <strong>of</strong> its range.<br />

Taxonomy and Description: The Romanian populations are<br />

identical to those <strong>of</strong> central Europe (Bálint and Szabó 1981).<br />

Distribution: Transpalearctic. The European populations are<br />

very scattered (Meyer 1981–1982). Only three sites are known<br />

in Transylvania (western part <strong>of</strong> Romania). One <strong>of</strong> them was<br />

reported at the beginning <strong>of</strong> this century and has never been<br />

confirmed. The second record originates from the 1970s and<br />

this, also, has not been confirmed recently. Only the third, the<br />

most recently discovered site (in 1979, Szabo 1982), seems to<br />

be strong enough to survive.<br />

Population Size and Status: Not known.


Habitat and Ecology: Hygrophilous, bivoltine (April, early<br />

May and July), stenotopic. Caterpillars myrmecoxenous (Fiedler<br />

1991). The ecology <strong>of</strong> the Transylvanian populations is not<br />

known.<br />

Threats: Wetland drainage, overcollecting.<br />

<strong>Conservation</strong>: The habitat <strong>of</strong> the strongest population should<br />

urgently be declared a nature reserve and the whole ecosystem<br />

also needs protection. A part <strong>of</strong> the swamp has been destroyed<br />

already, and Maculinea alcon L. has disappeared. The butterfly<br />

must be protected by law, because it is the subject <strong>of</strong> intensive<br />

collecting.<br />

Lycaena tityrus argentifex Bálint<br />

Country: Romania.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

Endemic to the alpine and subalpine zones <strong>of</strong> the eastern and<br />

southern Carpathians.<br />

Taxonomy and Description: Somewhat similar to the taxon<br />

subalpinus Speyer, but the female <strong>of</strong>ten with submarginal<br />

orange lunules, and the underside ground colour deep silver<br />

grey with large spots.<br />

Distribution: West Palaearctic species. The subspecies<br />

argentifex is confined to the eastern and southern Carpathians,<br />

occurring at the same elevations as subalpinus in the Alps<br />

(800–2000m).<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Alpicol, univoltine (July–August),<br />

presumed stenotopic. Caterpillar myrmecoxenous (Fiedler<br />

1991). The habitats are mesophilous meadows, but other details<br />

concerning the ecology <strong>of</strong> the subspecies are not known.<br />

Threats: Intensification <strong>of</strong> grassland management and wetland<br />

drainage. The known habitats are traditionally used by the<br />

native inhabitants as grassland. Any modification <strong>of</strong> the system<br />

by changes in the regime or the use <strong>of</strong> the meadows could be<br />

dangerous for the species.<br />

<strong>Conservation</strong>: Establishment <strong>of</strong> natural reserves and the study<br />

<strong>of</strong> the ecology <strong>of</strong> the species are required.<br />

110<br />

Pseudophilotes bavius hungaricus<br />

(Diószeghy)<br />

Country: Romania.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered.<br />

An endemic subspecies <strong>of</strong> the Transylvanian Basin (West<br />

Romania).<br />

Taxonomy and Description: Close to P. bavius Eversmann,<br />

but differing in some minor characters (König 1988).<br />

Distribution: An east Mediterranean species. The western<br />

limit <strong>of</strong> its range is represented by the populations <strong>of</strong> hungaricus.<br />

Only very few populations are known and they are very scattered<br />

(Szabó 1982). Very recently the species (most probably an<br />

undescribed subspecies) was found in Dobrogea, south Romania<br />

(Székely, pers. comm.).<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Xerothermophilous, univoltine (April),<br />

stenotopic. Caterpillar presumed weakly myrmecophilous<br />

(Fiedler 1991). Larval hostplant Salvia nutans L. The early<br />

stages were described and the ecology was partly studied by<br />

König (1988).<br />

Threats: Intensification <strong>of</strong> grassland management, afforestation,<br />

earthworks, air pollution and overcollecting. Some habitats<br />

were overgrazed (Cluj Napoca-Kolozsvárl: Szénafiivek), or<br />

planted over by Pinus nigra Arn. or Robinia pseudacacia L.<br />

(Teuis-Tövis). Others are endangered by a chemical factory<br />

(close to Sibiu-Nagyszeben). A single colony in a nature<br />

reserve (Suat-Magyarszovát: Csigla domb) was most probably<br />

exterminated by the activities <strong>of</strong> the native collectors.<br />

<strong>Conservation</strong>: Some habitats are protected by law, but the<br />

protection remains only on paper and has not been enforced.<br />

Very strict dispositions would be needed. Thorough ecological<br />

studies are necessary to continue the initial work <strong>of</strong> König<br />

(1988).<br />

Tomares nogelii dobrogensis (Caradja)<br />

Country: Romania.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered.<br />

Only a single population is known from Romania.<br />

Taxonomy and Description: see Higgins and Riley (1983)<br />

and Hesselbarth and Schurian (1984).


Distribution: An east Mediterranean species. The Romanian<br />

locality in Dobrogea represents the most westerly one known.<br />

Population Size and Status: Not known.<br />

Habitat and Ecology: Xerothermophilous, univoltine (May).<br />

Caterpillar steadily myrmecophilous (Fiedler 1991). Other<br />

details concerning the Romanian populations are not known.<br />

Threats: Overgrazing, overcollecting.<br />

<strong>Conservation</strong>: The known site is very small and strongly<br />

disturbed by human activity. The habitat is well known amongst<br />

the Romanian lepidopterists, so overcollecting <strong>of</strong> the species is<br />

a real danger in spite <strong>of</strong> the fact that the habitat is a nature<br />

reserve. The Romanian population <strong>of</strong> nogelii has never been<br />

studied from the ecological point <strong>of</strong> view, and field studies <strong>of</strong><br />

nogelii dobrogensis are urgently required.<br />

References<br />

BÁLINT, Zs. 1985a. Cupido osiris (Meigen, 1829) in the Carpathian Basin<br />

(Lepidoptera: <strong>Lycaenidae</strong>). Folia ent. hung, 46: 256–257 (in Hungarian).<br />

BÁLINT, Zs. 1985b. Plebicula dorylas magna nov. ssp. (Lep.: <strong>Lycaenidae</strong>)<br />

from the eastern Carpathians, Romania. Neue Ent. Nachr. 14: 14–20.<br />

BÁLINT, Zs. 1986. Further studies on Maculinea alcon (Den. & Schiff.,<br />

1775) (Lepidoptera: <strong>Lycaenidae</strong>). Galathea, Nürnberg 2: 92–108.<br />

BÁLINT, Zs. 1991. <strong>Conservation</strong> <strong>of</strong> butterflies in Hungary. Oedippus 3:<br />

5–36.<br />

BÁLINT, Zs. and SZABÓ, Gy. 1981. The occurrence <strong>of</strong> Lycaena helle (Den.<br />

et Schiff., 1775) in the Szatmár lowland. Folia ent. hung. 42: 235–236 (in<br />

Hungarian).<br />

111<br />

DE LESSE, H. 1960. Les nombres de chromosomes dans la classification du<br />

groupe d'Agrodiaetus ripartii Freyer (Lepidoplera, <strong>Lycaenidae</strong>). Rev.<br />

franc. Ent. 27: 240–264.<br />

FIEDLER, K. 1991. European and North West African <strong>Lycaenidae</strong><br />

(Lepidoptera) and their associations with ants. J. Lepid. Soc. 28:239–257.<br />

GEIGER, W. (Ed.) 1988. Tagfalter und ihre Lebensräume. Arten. Gefährdung.<br />

Schutz. Schweizerische Bund für Naturschutz, Basel, 2nd edn., xi + 516<br />

pp.<br />

HESSELBARTH, G. and SCHURIAN, K.G. 1984. Beitrag zur Taxonomie,<br />

Verbreitung und Biologie von Tomares nogelii (Herrich-Schäffer, 1851)<br />

in der Türkei (Lepidoptera, <strong>Lycaenidae</strong>). Entom<strong>of</strong>auna 5: 243–250.<br />

HIGGINS, L.G. and RILEY, N.D. 1983. A Field Guide to the <strong>Butterflies</strong> <strong>of</strong><br />

Britain and Europe. Collins, London, 5th edn., 384 pp.<br />

KÖNIG, F. 1988. Morphological, biological and ecological data on Philotes<br />

bavius hungarica Diószegtiy, 1913 (Lepidoptera, <strong>Lycaenidae</strong>). 4thNation.<br />

Conf. Entomol. Cluj-Napoca, pp. 175–182. (in Romanian).<br />

KRÁLICEK, M. and POVOLNŸ, D. 1957. Polyommatus eros eroides<br />

(Frivaldsky, 1837) (sic) in Czechoslovakia. Acta Soc. ent. Czech. 53:<br />

193–201 (in Czech.).<br />

KUDRNA, 0. 1986. Aspects <strong>of</strong> the <strong>Conservation</strong> <strong>of</strong> <strong>Butterflies</strong> in Europe.<br />

<strong>Butterflies</strong> <strong>of</strong> Europe. Volume 8. Aula-Verlag Wiesbaden, 323 pp.<br />

KULFAN, J. and KULFAN, M. 1991. Die Tagfalterfauna der Slowakei und<br />

ihr Schutz unter besonderer Berücksichtigung der Gebirgökosysteme.<br />

Oedippus 3: 75–102, 4 figs.<br />

MEYER, M. 1981–1982. Revision systematique, chorologique et ecologique<br />

des populations europeennes de Lycaena (Helleia) helle Denis &<br />

Schiffermüller, 1775.<br />

SZABÓ, A. 1982. Data to the Romanian distribution <strong>of</strong> the species Lycaena<br />

helle Schiff. and Philotes bavius Ev. (Lepidoptera, <strong>Lycaenidae</strong>). Studii si<br />

Comunicari, Reghin 2: 299–306 (in Romanian).<br />

SZABÓ, R. 1956. The Lycaenids <strong>of</strong> Hungary. Folia ent. hung. 9 (SN):<br />

235–262.<br />

UHRIK-MÉSZÂROS, 1948. Data on the knowledge <strong>of</strong> the life history <strong>of</strong><br />

Lycaena iolas O. Folia ent. hung. 3 (SN): 5–8 (in Hungarian).<br />

VARGA, Z. 1968. Bemerkungen und Ergänzungen zur taxonomishen<br />

Beurteilung und Ökologie der im Karpatenbecken vorkommenden<br />

Populationen von Ariciaartaxerxes Fabr.Acta biol. Debrecina 6:171–185.


Country: Japan.<br />

'Chosen-aka-shijimP, Coreana raphaelis Oberthur<br />

T. HIROWATARI<br />

Entomological Laboratory, College <strong>of</strong> Agriculture, University <strong>of</strong> Osaka Prefecture, Sakai, Osaka 591, Japan<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

In Japan, this species occurs only in very local areas <strong>of</strong><br />

northern Honshu, and all the populations seem to be threatened<br />

with extinction. In most districts, positive actions by the residents<br />

in support <strong>of</strong> conservation are evident. This is one <strong>of</strong> the model<br />

cases <strong>of</strong> conservation in Japan (Sibatani 1989), and activities<br />

have included: (a) monitoring population size; (b) volunteer<br />

patrolling; (c) breeding <strong>of</strong> butterflies; (d) planting <strong>of</strong> foodplants;<br />

and (e) campaigns for public education in conservation<br />

awareness.<br />

Taxonomy and Description: The genus Coreana Tutt is<br />

monotypic and one <strong>of</strong> the most primitive members <strong>of</strong> the tribe<br />

Theclini (sensu Eliot 1973). Representatives from Iwate<br />

Prefecture are treated as ssp. yamamotoi Okano (broader orange<br />

marking on upperside), and from Yamagata Prefecture as ssp.<br />

ohruii Shirozu (broader black border on hindwing upperside).<br />

Distribution: Confined to northern Honshu Iwate Pref.: Kuji<br />

(40°ll'N, 141°46'E) to Miyako (39°38'N, 141°59'E), around<br />

Sizukuishi (39°41'N, 140°59'E); Yamagata Pref.: Shinjo<br />

(38°45'N, 140°18'E) to Kawanishi (37°59'N, 140°02'E);<br />

Niigata Pref.:Asahi(38°15'N,139°35'E),Sekikawa(36°06'N,<br />

139°36'E).<br />

Outside Japan, this species is distributed in some disjunct<br />

areas <strong>of</strong> Amur and southern Ussuri (southeastern Russia),<br />

northern China and the Korean Peninsula.<br />

Habitat and Ecology: In northern Honshu the species occurs<br />

on the borders <strong>of</strong> woods, along creeks, in marshes near low hills<br />

where Fraxinus spp. (Oleaceae) grows, the foodplant <strong>of</strong> this<br />

species. There are many rural houses in this area which are<br />

surrounded by hedges <strong>of</strong> this plant, and such hedges are now the<br />

main habitat <strong>of</strong> this species.<br />

The larvae mainly feed on Fraxinus japonica, and<br />

occasionally on F. lanuginosa and F. mandshurica. Eggs are<br />

laid in batches <strong>of</strong> several to 20 on the trunks <strong>of</strong> foodplants.<br />

Larvae hatch in April. Second instar larvae make simple shelters<br />

by spinning leaves. Third and 4th instar larvae nibble the stems<br />

112<br />

<strong>of</strong> the leaves and sit on drooping leaves. Fourth (last) instar<br />

larvae are sometimes tended by ants (Lasius spp.). Pupae are<br />

found under fallen leaves, pieces <strong>of</strong> decayed wood or under<br />

stones (Fukuda et al. 1984).<br />

There is one generation each year. Adults emerge in mid-<br />

June, and are seen until early August. The species overwinters<br />

as eggs.<br />

Threats: Habitats <strong>of</strong> this species are being progressively<br />

destroyed by the development <strong>of</strong> housing sites, etc. Felling<br />

foodplants in the area along creeks or marshes near low hills is<br />

one <strong>of</strong> the main factors which reduces the population size. One<br />

further threat is the decrease in the number <strong>of</strong> hedges made <strong>of</strong><br />

Fraxinus spp. which now seem to be essential for the species.<br />

<strong>Conservation</strong>: This species has been strongly associated with<br />

rural human life in its range in Japan and its survival seems to<br />

depend on the future activities <strong>of</strong> the local communities.<br />

In Kawanishi (Yamagata Pref.), this species was designated<br />

as a protected species <strong>of</strong> the Prefecture in 1977. However, in the<br />

absence <strong>of</strong> effective conservation measures, its habitats have<br />

been progressively destroyed. Nevertheless, local bodies began<br />

to promote conservation and urged the local government to take<br />

conservation measures. For example, in 1989, at the time <strong>of</strong><br />

creek improvement in Kawanishi, the foodplants <strong>of</strong> C. raphaelis<br />

were transplanted to unoccupied ground (c. 600m) that had<br />

been taken over by the local government, and some <strong>of</strong> the plants<br />

were also transplanted to school grounds (Nagaoka amd Kusakari<br />

1989).<br />

In Iwate Pref., the butterfly is now designated a protected<br />

species in all the cities, towns and villages (Ogata et al. 1989).<br />

As in the case <strong>of</strong> Yamagata Pref., local governments as well as<br />

local bodies do recognize that prohibition <strong>of</strong> collecting without<br />

adequate management measures are not effective for real<br />

protection <strong>of</strong> butterflies, and they are now devising the following<br />

measures: (1) continuous monitoring <strong>of</strong> population size by<br />

local bodies and lepidopterists; (2) environmental protection by<br />

patrolling and by checking the urbanisation plans <strong>of</strong> the local<br />

governments; (3) breeding butterflies and planting <strong>of</strong> their<br />

foodplants in public places, such as school grounds or private<br />

housing sites; and (4) education for conservation awareness by


holding lecture meetings and by producing public relations<br />

leaflets or videotapes.<br />

References<br />

FUKUDA, H., HAMA, E., KUZUYA, T., TAKAHASHI, A., TAKAHASHI,<br />

M., TANAKA, B., TANAKA, H., WAKABAYASHI, M. and<br />

WATANABE, Y. 1984. The Life Histories <strong>of</strong> <strong>Butterflies</strong> in Japan Vol. 3.<br />

Hoikusha, Osaka (in Japanese).<br />

NAGAOKA, H. and K. KUSAKARI. 1989. Habitat status <strong>of</strong> Coreana raphaelis<br />

113<br />

and conservation practice in Kawanisi-shi, Yamagata Prefecture.<br />

In: Hama, E., Ishii, M. and Sibatani, A. (Eds) Decline and<br />

<strong>Conservation</strong> <strong>of</strong> <strong>Butterflies</strong> in Japan, I. pp. 65–73.<br />

Lepidopterological Society <strong>of</strong> Japan, Osaka (in Japanese).<br />

OGATA, Y., MIURA, H. and Y. MUROYA. 1989. Population status<br />

<strong>of</strong> Coreana raphaelis and the conservation practices in Iwate<br />

Prefecture. In: Hama, E., Ishii, M. and Sibatani, A. (Eds) Decline<br />

and <strong>Conservation</strong> <strong>of</strong> <strong>Butterflies</strong> in Japan, I. pp. 74–82 (in Japanese).<br />

SIBATANI, A. 1989. Decline and conservation <strong>of</strong> butterflies in Japan.<br />

In: Hama, E., Ishii, M. and Sibatani, A. (Eds) Decline and<br />

<strong>Conservation</strong> <strong>of</strong> <strong>Butterflies</strong> in Japan 1, pp. 16–22.<br />

Lepidopterological Society <strong>of</strong> Japan, Osaka.


Country: Japan.<br />

'Oruri-shijimi', Shijimiaeoides divinus Fixsen<br />

T. HIROWATARI<br />

Entomological Laboratory, College <strong>of</strong> Agriculture, University <strong>of</strong> Osaka Prefecture, Sakai, Osaka 591, Japan<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

In Japan, this species occurs in very local and disjunct areas<br />

in Honshu and Kyushu. The populations <strong>of</strong> Honshu are feared<br />

extinct, especially in the northern district. In fact, the ones in<br />

Aomori Prefecture and Iwate Prefecture are thought to have<br />

been extinct since the late 1970s. No effective conservation<br />

measures have been taken by local governments in Kyushu<br />

except for the prohibition <strong>of</strong> collecting.<br />

Taxonomy and Description: The genus Shijimiaeoides Beuret,<br />

related to Maculinea van Eecke and Glaucopsyche Scudder,<br />

contains two species, lanty Oberthur and divinus Fixsen, both<br />

occurring in eastern Asia. Representatives from Honshu and<br />

Kyushu are treated as ssp. barine Leech and asonis Matsumura,<br />

respectively <strong>of</strong> S. divinus. The nominotypical subspecies occurs<br />

in the Korean Peninsula, and also seems to be endangered<br />

(Fukuda et al. 1984).<br />

Distribution: Very local, in disjunct areas <strong>of</strong> Honshu and<br />

Kyushu. The localities <strong>of</strong> known habitats consist <strong>of</strong> three<br />

groups: (1) northern Honshu: Aomori Pref., Iwate Pref. Morioka<br />

(39°41 'N, 141 °08'E); (2) central Honshu: Niigata, Nagano and<br />

Gunma Prefs.; and (3) Kyushu: Aso, Kuju. Outside Japan, this<br />

species is distributed in local areas <strong>of</strong> the Korean Peninsula.<br />

Population Size: Population size <strong>of</strong> this species has never been<br />

estimated systematically. However, the decline <strong>of</strong> populations<br />

in northern Honshu seems to be obvious. In Aomori Pref. they<br />

began to decrease in the 1960s and became extinct by the<br />

second half <strong>of</strong> the 1970s (Muroya 1989). In Azumino, Nagano<br />

Pref., this species was widespread in the 1970s, but it is now on<br />

the brink <strong>of</strong> extinction (Kobayashi 1989). Exceptionally in<br />

Kyushu, populations <strong>of</strong> this species and their habitats seem to<br />

have been conserved during these decades.<br />

Habitat and Ecology: Usually occurs on sunny grassland,<br />

river banks, degraded areas along railroads and, especially, on<br />

volcanic slopes.<br />

114<br />

Larvae obligately feed on the flowers and flowerbuds <strong>of</strong><br />

Sophora flavescens (Leguminosae). In Kyushu (Aso, Kuju),<br />

volcanic slopes are used for pasture, and S. flavescens grows<br />

well in such grasslands, because cattle avoid grazing it.<br />

There is one generation each year, with adults present in<br />

May in Kyushu, and in June in Honshu. Eggs are laid on a<br />

flowerbud. The larva is yellowish and milky white, attended by<br />

ants (not identified). The pupa is blackish and found on the<br />

ground near the foodplant.<br />

Threats: Extinction <strong>of</strong> the northern Honshu population (Aomori,<br />

Iwate) is attributable to degradation <strong>of</strong> habitat caused by<br />

cultivation, alteration in land management practices, and<br />

urbanisation such as construction <strong>of</strong> golf courses, airport, and<br />

so on. The habitats <strong>of</strong> S. divinus are mostly grasslands, which<br />

may be easily cultivated or used in urban development. For<br />

example, at the foot <strong>of</strong> Mt. Iwaki (Aomori Pref.), exploitation<br />

by the national and local governments was carried out for seven<br />

years (1962–1968) during which time a total <strong>of</strong> 2155ha <strong>of</strong><br />

grassland were cultivated, covering most <strong>of</strong> the habitats <strong>of</strong> the<br />

species. After that small populations that had survived in<br />

grasslands along creeks also became extinct in the late 1970s.<br />

Alterations in land management practices, another <strong>of</strong> the<br />

main factors, has involved a decrease in the number <strong>of</strong> cattle<br />

and a resultant cessation <strong>of</strong> grass-cutting: this has led to<br />

successional changes which are not suitable for the growth <strong>of</strong><br />

the foodplant.<br />

<strong>Conservation</strong>: In spite <strong>of</strong> the decline in the 1960s in northern<br />

Honshu, no campaign was undertaken to save the species.<br />

Consequently, the populations <strong>of</strong> Aomori were eradicated in<br />

the late 1970s. The populations <strong>of</strong> central Honshu seem to be at<br />

a crucial stage at the present time but no conservation project<br />

has been undertaken, either. In Kyushu, the local governments<br />

prohibit collecting <strong>of</strong> the species, and local bodies are cooperating<br />

on a volunteer patrolling system. An essential measure would<br />

be to conserve the grassland and prevent any further alterations<br />

in land management practices. However, at this moment nothing<br />

has been done as a result <strong>of</strong> inadequate financial support.


References<br />

FUKUDA, H., HAMA, E., KUZUYA, T., TAKAHASHI, A., TAKAHASHI,<br />

M., TANAKA, B., WAKABAYASHI, M. and WATANABE, Y. 1984.<br />

The Life Histories <strong>of</strong> <strong>Butterflies</strong> in Japan, Vol. 3. Hoikusha, Osaka (in<br />

Japanese).<br />

115<br />

KOBAYASHI, Y. 1989. Decline <strong>of</strong> Shijimiaeoides divinus populations in<br />

Azumino, Nagano Prefecture, pp. 97–98, in Hama, E., Ishii, M. and<br />

Sibatani, A. (Eds) Decline and <strong>Conservation</strong> <strong>of</strong> <strong>Butterflies</strong> in Japan I.<br />

Lepidopterological Society <strong>of</strong> Japan, Osaka (in Japanese).<br />

MUROYA, Y. 1989. Decline <strong>of</strong>Shijimiaeoides divinus populations in Aomori<br />

Prefecture. Ibid.: 90–97 (in Japanese).


Lange's Metalmark, Apodemia mormo langei Comstock<br />

Jerry A. POWELL 1 and Michael W. PARKER 2<br />

1 Department <strong>of</strong> Entomological Sciences, 201 Wellman Hall, University <strong>of</strong> California, Berkeley,<br />

California 94720-0001, U.S.A.<br />

2 U.S. Fish and Wildlife Service, San Francisco Bay National Wildlife Refuge Complex, Newark, CA 94560, U.S.A.<br />

Country: U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – possibly out <strong>of</strong><br />

danger; endangered listing (USFWS).<br />

This local subspecies was one <strong>of</strong> the first eight insects to be<br />

listed as an endangered species in 1976 under the Federal<br />

Endangered Species Act. Its remnant habitat was purchased by<br />

the U.S. Fish and Wildlife Service (USFWS) in 1980 and<br />

designated as the Antioch Dunes National Wildlife Refuge, the<br />

first such refuge in the country established to protect insects and<br />

plants. Population numbers <strong>of</strong> Apodemia mormo langei declined<br />

over a period <strong>of</strong> 50 years to a few hundred individuals in the<br />

early 1980s. During the past decade they have increased tenfold<br />

or more in response to recovery actions, which include exclusion<br />

<strong>of</strong> vehicular and most foot traffic, removal <strong>of</strong> exotic vegetation,<br />

and extensive outplanting <strong>of</strong> the larval host plant.<br />

General accounts <strong>of</strong> the unique insect and plant communities<br />

and destruction <strong>of</strong> the Antioch dunes by sand mining and<br />

Female Lange's Metalmark on flowers <strong>of</strong> Senecio douglasii.<br />

116<br />

industrial development have been recorded (Howard 1983;<br />

Howard and Arnold 1980; Farb 1964; Powell 1981, 1983).<br />

Taxonomy and Description: Considerable polytypy is<br />

expressed, <strong>of</strong>ten discordantly, in hostplant species, seasonal<br />

phenology and voltinism, and in size and colour pattern <strong>of</strong> the<br />

adults (Opler and Powell 1962; Powell 1975).<br />

Distribution: Apodemia mormo (Felder & Felder) (Riodininae)<br />

is widely distributed in the western Nearctic and occurs in<br />

scattered, <strong>of</strong>ten isolated colonies. Some forms <strong>of</strong> the butterfly<br />

are quite limited in geographic distribution and may be separated<br />

by distances <strong>of</strong> only 15–20km from other populations that are<br />

easily distinguished phenotypically. Apodemia mormo langei<br />

was discovered in 1933, on the riverine sand dune system just<br />

east <strong>of</strong> Antioch, Contra Costa Co., California (Comstock 1938).<br />

This subspecies is represented by just one population (Opler<br />

and Powell 1962). The nearest known colonies <strong>of</strong> Apodemia<br />

mormo are located on the northwest slope <strong>of</strong> Mt. Diablo and on<br />

the hills northwest <strong>of</strong> Vallejo, Solano Co., which are respectively<br />

about 15km SW and 39km WNW <strong>of</strong> the langei site. These<br />

populations have appreciable differences in colour pattern from<br />

langei.<br />

Population Size: U.S. Geological Survey topographical maps<br />

and aerial photographs that span 1905 to 1969, document the<br />

destruction <strong>of</strong> the Antioch sand dunes. At the turn <strong>of</strong> the<br />

century, after agricultural development <strong>of</strong> the region, they<br />

extended for 3–4km in a narrow band along the San Joaquin<br />

River and reached heights <strong>of</strong> 20–35m. Sand mining operations<br />

began prior to and during the 1920s (Howard and Arnold 1980).<br />

Although the installation <strong>of</strong> two powerline towers, in 1909 and<br />

1927, resulted in habitat disturbance including building<br />

construction and introduction <strong>of</strong> exotic plants, today these<br />

Pacific Gas and Electric (P.G. & E.) towers stand on the two<br />

remaining remnants <strong>of</strong> sand hills.<br />

By 1931, five huge sand pits, each with a railroad spur, were<br />

in operation, for the developing San Francisco Bay area. During<br />

post World War II years, massive industrial developments<br />

replaced the eastern half <strong>of</strong> the dunes, and in the 1950s sand<br />

mining moved to the western sector. A Kaiser Gypsum plant


was completed in 1956, and this isolated the two remaining<br />

remnants <strong>of</strong> the habitat, the Stamm Property (SP) to the west<br />

and the 'Little Corral' (LC) flanked by the P.G. & E. properties<br />

to the east (Figure 1).<br />

Although its habitat was gradually restricted during the 40<br />

years following discovery, the local abundance <strong>of</strong> Apodemia<br />

mormo langei remained high. Lepidopterists observed the<br />

butterflies by the hundreds and collected specimens at rates <strong>of</strong><br />

25–30 per hour in various seasons during 1947–1972. Even as<br />

late as 1972, three observers took 70 specimens one day and 50<br />

more seven days later and estimated sighting 150–200 on each<br />

date.<br />

Soon thereafter, however, the colonies were severely affected<br />

by increased sand-mining at the western parcel (SP), by<br />

rototilling on the P.G. & E. properties in compliance with a<br />

county ordinance for fire prevention, and by overgrazing by<br />

horses. Thus by 1976, when Powell began population census by<br />

transect counts, numbers <strong>of</strong> langei had dropped alarmingly.<br />

During mark-recapture monitoring by Arnold and Powell from<br />

1977–1982, the maximum number <strong>of</strong> langei two persons could<br />

observe in 3.5–4.5 hour periods at the two sites was 45–55<br />

(Arnold and Powell 1983; Powell unpubl. data). Total population<br />

numbers, calculated from daily Jolly-Seber and Manly-Parr<br />

estimates and the survival rates <strong>of</strong> individual butterflies, declined<br />

from more than 2000 in 1977 to fewer than 600in 1982 (Arnold<br />

1985 and unpubl. reports to USFWS).<br />

The butterfly numbers stabilised and increased slightly in<br />

1983–84 and then rose significantly as langei began to occupy<br />

Eriogonum in peripheral areas that had been both planted and<br />

colonised naturally. The estimated total population exceeded<br />

1200 in 1985, the final year <strong>of</strong> Arnold's mark-recapture studies<br />

(Arnold, unpublished report to USFWS). Subsequently, single<br />

day sighting counts, made weekly by USFWS personnel, indicate<br />

that the numbers <strong>of</strong> langei have continued to climb dramatically.<br />

Seasonal peak numbers have risen from 168 in 1986 to more<br />

than 1900 in 1991. It is likely that the population has 10 to 20<br />

times the number <strong>of</strong> butterflies that it was estimated to have<br />

contained 10 years ago (e.g. 6000 to 12,000).<br />

Habitat and Ecology: Populations occur in close association<br />

with the larval foodplants, species <strong>of</strong> Eriogonum (Polygonaceae),<br />

typically in well drained semiarid sites, such as rocky desert<br />

slopes, sand dunes, or chaparral-covered hills, ranging from sea<br />

coast to timberline at 2750m. Colonies <strong>of</strong> A. m. langei are<br />

limited to dense or moderately dense patches <strong>of</strong> the larval<br />

foodplant, Eriogonum nudum auriculatum (Arnold and Powell<br />

1983). Arnold and Powell believed that isolated or spindly,<br />

scattered plants fail to support colonies <strong>of</strong> the butterfly because<br />

early instar larvae derive insufficient protection for<br />

overwintering.<br />

A. m. langei is univoltine, with adults flying for about 30<br />

days beginning in early August. Males precede females by a<br />

few days, and peak numbers occur about two weeks after<br />

emergence begins. The eggs are deposited in clusters <strong>of</strong> 2–4 on<br />

withering foliage on the lower half <strong>of</strong> the plant. Eclosion <strong>of</strong><br />

larvae takes place during winter, after rainfall and foliation<br />

begin. Feeding occurs by skeletonizing the leaf surfaces and the<br />

inflorescence stalks by later instars in June and July. Larvae<br />

feed in early morning and presumably evening and retreat to the<br />

base <strong>of</strong> the plants during the day. Pupation occurs in the litter<br />

Figure 1. Map <strong>of</strong> the Antioch Dunes National Wildlife Refuge, just east <strong>of</strong> Antioch, Contra Costa Co., California.<br />

Redrawn from Antioch North Quadrangle, U.S. Geological Survey topographic map, 1953. Shaded areas delineated by dotted lines indicate construction sites,<br />

1953–1968, compiled from aerial photographs. Sparsely doited areas depict sand-mining excavations during 1953–1967. The two corridors traversed by P.G.<br />

& E. powerlines are the only unexcavated hills that remain. Bold lines define the two parcels <strong>of</strong> the Refuge, "Stamm Property" (SP) and "Little Corral" (LC),<br />

the latter flanked by P.G. & E. properties.<br />

117


at the base <strong>of</strong> the plant, in late July and early August<br />

(Arnold and Powell 1983).<br />

Males tend to occupy restricted areas day after day,<br />

while females move greater distances. Mark-recapture<br />

data suggested that males live an average <strong>of</strong> about 12<br />

days. Adults <strong>of</strong> both sexes forage for nectar; Eriogonum<br />

serves as the primary nectar source, at least in recent<br />

years, while occasional visits are made to Gutierrezia<br />

and Senecio (Asteracecae), both <strong>of</strong> which were formerly<br />

more abundant, and other plants (Arnold and Powell<br />

1983).<br />

Under natural circumstances, this variety <strong>of</strong><br />

Eriogonum nudum probably lived as an edge species,<br />

occupying slip slopes <strong>of</strong> active sand in the hills that were<br />

stabilised by scattered oaks and a rich flora <strong>of</strong> desert<br />

affinities (Howard 1983). There is no doubt that the<br />

active sand habitat greatly increased during the 1920–1940<br />

era <strong>of</strong> sand-mining. Thus, it is likely that, along with its<br />

larval host, A. m. langei had increased in population<br />

numbers by the time <strong>of</strong> its discovery in the early 1930s.<br />

<strong>Conservation</strong>: Rehabilitation <strong>of</strong> habitat suitable for<br />

colonisation by A. m. langei begore before acquisition <strong>of</strong><br />

the Refuge lands by USFWS. Propagation and outplanting<br />

on the P.G. & E. properties began in 1979, following<br />

inadvertent rototilling <strong>of</strong> one <strong>of</strong> the primary stands <strong>of</strong><br />

Eriogonum, despite efforts by the company to prevent<br />

such an accident. P.G. & E. contracted with Biosystems<br />

Analysis Inc. to develop a restoration plan, and about 450<br />

seedlings were planted in March 1980 (Howard and<br />

Arnold 1980). Collective efforts during 1980–1984, which<br />

were orchestrated by Howard and Arnold, were financed<br />

by P.G. & E., grants from the California Native Plant<br />

Society, a USFWS contract, and assistance by the<br />

University <strong>of</strong> California, Berkeley undergraduate botany<br />

association and other volunteers (Arnold 1985). During<br />

the same years, the USFWS began planting Eriogonum at<br />

the western parcel (SP) by scarifying and seeding the<br />

excavated surface left bare by a final surge <strong>of</strong> sandmining<br />

in 1978–79. In 1985 the USFWS entered a<br />

Cooperative Agreement with P.G. & E. that allows the<br />

Service to manage the additional 5ha owned by P.G. &<br />

E., which are contiguous with LC (Figure 1). In 1987,<br />

fencing <strong>of</strong> the lands was completed, and this virtually<br />

eliminated further human degradation <strong>of</strong> the habitat.<br />

Also in 1987 a vineyard to the south <strong>of</strong> the Apodemia<br />

colony at SP was removed and subsequently planted with<br />

7000 Eriogonum seedlings. Small numbers <strong>of</strong> langei<br />

have begun to occupy that area (58 were observed in one<br />

day in 1991).<br />

In 1991 a more ambitious cooperative restoration<br />

project was initiated to create new sand dunes in previously<br />

mined areas occupied by weedy vegetation. Low hills,<br />

consisting <strong>of</strong> sand mined from another P.G. & E. property<br />

up river, have been deposited and contoured on the LC<br />

and P.G. & E. western parcels. These were subsequently<br />

118<br />

seeded and planted and now bear mantles <strong>of</strong> Eriogonum<br />

and Senecio douglasii seedlings, the latter a major nectar<br />

source for the butterflies, as well as two endangered<br />

plants, Oenothera deltoides var. howelli (Onagraceae)<br />

and Erysimum capitatum var. angustatum (Brassicaceae).<br />

Altogether, we estimate that outplanted Eriogonum<br />

colonies that have reached densities believed to be<br />

sufficient to support Apodemia occupy areas <strong>of</strong> 9600m 2<br />

at SP and 3200m 2 on the P.G. & E. properties, a total <strong>of</strong><br />

nearly 1.3ha.<br />

In 1979 the maximum distribution <strong>of</strong> the foodplant<br />

range (range width x range length, right angle) was<br />

estimated to be 2.3ha at LC and 1.5ha at SP, not more than<br />

one-third <strong>of</strong> which was suitable for Apodemia as judged<br />

by occurrence <strong>of</strong> adults (Arnold and Powell 1983). Hence,<br />

there was only about 1.3ha <strong>of</strong> viable Eriogonum habitat<br />

at its lowest ebb, and subsequent efforts have at least<br />

doubled that area. Several additional patches <strong>of</strong> the host<br />

plant, planted in recent years, are expected to develop<br />

into viable habitat within a few seasons.<br />

Threats: Effects <strong>of</strong> weediness and possibility <strong>of</strong> fire are<br />

the principal threats to continued existence <strong>of</strong> A. m.<br />

langei. A standing crop <strong>of</strong> annual weeds poses a fire<br />

danger each year during the long dry season<br />

(May-October). The refuge is bordered inland by<br />

industrial development, and there are several small<br />

beaches that are accessible to recreational boat visits, so<br />

possible sources <strong>of</strong> human-initiated fires cannot be<br />

controlled.<br />

<strong>Conservation</strong> and recovery efforts have been a<br />

dramatic success, increasing the population numbers<br />

from a perilously low level a decade ago. However, after<br />

10–15 years Eriogonum host plants senesce, and they fail<br />

to reproduce in the absence <strong>of</strong> open, active sand. We have<br />

witnessed the growth and decline <strong>of</strong> Eriogonum and<br />

associated langei colonies in several areas during the past<br />

15 years. For example, a robust colony developed in<br />

association with the foundations <strong>of</strong> a building that was<br />

removed after 1972 on P.G. & E. west; it was the home<br />

<strong>of</strong> a strong colony <strong>of</strong> langei during 1978–1982, but these<br />

plants were senescent by 1988 and have died out by 1992.<br />

The foundations protected the plants from rototilling<br />

during the 1970s, but competition with weeds prevented<br />

seedling growth. A similar fate can be predicted for most<br />

<strong>of</strong> the existing and recently planted patches <strong>of</strong> Eriogonum.<br />

Once a colony <strong>of</strong> the food plant is established, it is<br />

impractical to prevent a ground cover <strong>of</strong> weeds, especially<br />

annual exotic grasses, yellow star thistle, Russian thistle,<br />

and vetch, which does not kill mature Eriogonum but<br />

prevents development <strong>of</strong> its seedlings. Thus, it is easier<br />

to clear a site and plant new patches <strong>of</strong> the host plant than<br />

to maintain existing ones.<br />

To be successful on a long term basis, the management<br />

plan needs to prescribe replacing senescent patches <strong>of</strong><br />

Eriogonum on a continuing basis.


Acknowledgements<br />

We thank Richard Arnold, Alice Howard, John Steiner, and<br />

Jean Takekawa for discussions and information regarding the<br />

recent and past history <strong>of</strong> the Antioch sand dunes and the flora;<br />

Arnold and Steiner reviewed a draft <strong>of</strong> the manuscript and<br />

<strong>of</strong>fered useful criticisms.<br />

References<br />

ARNOLD, R.A. 1985. Private and government-funded conservation programs<br />

for endangered insects in California. Nat. Areas J. 5: 28–39.<br />

ARNOLD, R.A. and POWELL, J.A. 1983. Apodemia mormo langei. pp.<br />

99–128. In: Arnold, R. A., Ecological studies <strong>of</strong> six endangered butterflies<br />

(Lepidoptera: <strong>Lycaenidae</strong>): Island biogeography, patch dynamics, and<br />

the design <strong>of</strong> habitats preserves. Univ. Calif. Publns Entomol. 99.<br />

COMSTOCK, J.A. 1938. A new Apodemia from California. Bull. South Calif.<br />

119<br />

Acad. Sci. 37: 129–132.<br />

FARB, P. 1964. Insect city in the dunes, pp. 44–49. In: The Land and Wildlife<br />

<strong>of</strong> North America. Life Nature Library, Time-Life Books, New York.<br />

HOWARD, A.Q. (Ed.) 1983. The Antioch dunes. A report to the U.S. Fish and<br />

Wildlife Service. Prepared under P.O. No. 11640-0333-1; 115 pp. +<br />

appendices.<br />

HOWARD, A.Q. and ARNOLD, R.A. 1980. The Antioch dunes – safe at last?<br />

Fremontia 8(3): 3–12.<br />

OPLER, P.A. and POWELL, J.A. 1962. Taxonomic and distributional studies<br />

on the western components <strong>of</strong> the Apodemia mormo complex (Riodinidae).<br />

J. Lepid. Soc. 15: 145–171.<br />

POWELL, J.A. 1975. Riodinidae. The Metalmarks. pp. 259–272. In: Howe,<br />

W.H. (Ed.). The <strong>Butterflies</strong> <strong>of</strong> North America. Doubleday; Garden City,<br />

New York.<br />

POWELL, J.A. 1981. Endangered habitats for insects: California coastal sand<br />

dunes. Atala 6: 41–55 (1978).<br />

POWELL, J.A. 1983. Changes in the insect fauna <strong>of</strong> a deteriorating riverine<br />

sand dune community during 50 years <strong>of</strong> human exploitation. Report to<br />

U.S. Fish and Wildlife Service. Prepared under P.O. no. 11640-0333-1;<br />

78 pp.


The Hermes Copper, Lycaena hermes (Edwards)<br />

D.K. FAULKNER and J.W. BROWN<br />

Entomology Department, San Diego Natural History Museum, P.O. Box 1390, San Diego, California 92112, U.S.A.<br />

Country: U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status–rare; indeterminate<br />

(Red List).<br />

The Hermes Copper is a remarkably distinct species, differing<br />

considerably from its congeners in both morphology and<br />

ecology. It has a highly restricted geographical distribution in<br />

the southwestern United States and adjacent northwestern Baja<br />

California, Mexico. Because <strong>of</strong> the widespread loss and<br />

fragmentation <strong>of</strong> its habitats, associated with urbanisation and<br />

other development, this species has lost a significant portion <strong>of</strong><br />

its former range. Owing to its vulnerable nature, the Hermes<br />

Copper is recognized by the United States Fish and Wildlife<br />

Service as a 'category 2 candidate' species for listing as<br />

threatened or endangered.<br />

Taxonomy and Description: The Hermes Copper was described<br />

as Chrysophanus hermes by W.H. Edwards (1870) from<br />

' California'. Wright (1906) later described Chrysophanus delsud<br />

from San Diego County, California; the latter is a subjective<br />

synonym <strong>of</strong> hermes. The species has been included in the<br />

genera Tharsalea Scudder (e.g. Comstock 1927; Wright 1930)<br />

and Lycaena Fabricius (dos Passos 1964; Howe 1975). Miller<br />

and Brown (1979) placed hermes in the monotypic genus<br />

Hermelycaena Miller and Brown on the basis <strong>of</strong> its unique<br />

morphological and ecological characteristics. Currently, most<br />

taxonomists consider hermes to be a member <strong>of</strong> the Holarctic<br />

genus Lycaena.<br />

Distribution: The Hermes Copper is a narrowly endemic<br />

species, restricted to western San Diego County, California,<br />

and a small portion <strong>of</strong> adjacent Baja California, Mexico (Brown<br />

1980; Ehrlich and Ehrlich 1961; Emmel and Emmel 1973;<br />

Garth and Tilden 1986; Orsak 1977; Rindge 1948; Scott 1988).<br />

Its total range is approximately 250km from north to south<br />

(from about Fallbrook in San Diego County, California, to<br />

slightly south <strong>of</strong> Santa Tomas in Baja California, Mexico), and<br />

50km from east to west (from near the coast, inland to about<br />

Pine Valley, California). Within this range it occurs in small,<br />

disjunct colonies.<br />

120<br />

In San Diego County, the Hermes Copper has been recorded<br />

from El Cajon, Santee, Flynn Springs, Blossom Valley, Tecate,<br />

Pine Valley, Guatay, and numerous other localities, many <strong>of</strong><br />

which no longer support native vegetation.<br />

Population Size: In the absence <strong>of</strong> focused studies on population<br />

size and vagility, information on these parameters is mostly<br />

anecdotal. The Hermes Copper has been collected at about 35<br />

localities in the United States and four localities in Mexico.<br />

Most colonies are isolated from each other. Hence, gene flow<br />

between populations probably is rather low. In addition, adults<br />

exhibit limited vagility: they do not hilltop and they seldom are<br />

observed beyond the immediate vicinity <strong>of</strong> the larval host.<br />

Thorne (1963) indicated that colonies vary little in size from<br />

year to year, but there are few quantitative data to support this<br />

observation. It is likely that few colonies exceed 50 individuals.<br />

Habitat and Ecology: This species occurs in coastal sage scrub<br />

and open southern mixed chaparral communities in which the<br />

larval host plant, redberry (Rhamnus crocea Nutt. in T. & G.,<br />

Rhamnaceae) is a common component. In San Diego County,<br />

these habitat types range from near sea level along the coast to<br />

about 1250m at the western edge <strong>of</strong> the Laguna Mountains. The<br />

foodplant is a fairly common species, extending well beyond<br />

the range <strong>of</strong> the butterfly. Hence, the restricted distribution <strong>of</strong><br />

the Hermes Copper is difficult to explain.<br />

The Hermes Copper is univoltine, with adults present from<br />

mid-May through early July, depending upon elevation. The<br />

primary adult nectar source at most localities is flat-top<br />

buckwheat (Eriogonumfasciculatum Bentham; Polygonaceae),<br />

but L. hermes also has been observed to nectar on slender<br />

sunflower (Helianthus gracilentus) and a few other composites<br />

(Asteraceae). Males perch on vegetation along trails and<br />

openings, and confront other butterflies that pass by.<br />

Eggs are laid singly on stems <strong>of</strong> the foodplant and overwinter<br />

until the following spring. The egg is white, echinoid, and<br />

covered with deep pits between high, irregular walls. The fullygrown<br />

larva is apple green, with a mid-dorsal band <strong>of</strong> darker<br />

green bordered with yellowish-green. Pupation occurs on the<br />

foodplant, and the pupa is attached by a cremaster and a silken


girdle. Full details <strong>of</strong> the life history are presented by Comstock<br />

and Dammers (1935). Ballmer and Pratt (1989) indicate that the<br />

larvae <strong>of</strong> Lycaena hermes differ greatly from those <strong>of</strong> other<br />

Lycaena species in host preference and morphology. No<br />

parasitoids or predators are recorded.<br />

Threats: Although declines have not been documented<br />

quantitatively, the Hermes Copper undoubtedly has suffered<br />

from loss and fragmentation <strong>of</strong> habitat as a result <strong>of</strong> urbanisation.<br />

As long ago as 1930, Wright (1930) reported 'Its trysting places<br />

are being rapidly taken over by realtors and the species may<br />

soon become extinct...' Indeed, much <strong>of</strong> the former range <strong>of</strong> L.<br />

hermes is presently occupied by urbanised portions <strong>of</strong> San<br />

Diego. As development in San Diego County extends eastward<br />

from the coast, the Hermes Copper is further threatened by<br />

habitat loss.<br />

<strong>Conservation</strong>: Currently, the Hermes Copper receives minimal<br />

protection under the California Environmental Quality Act.<br />

Under this legislation, impacts to sensitive plants and animals<br />

and sensitive habitat types must be assessed to determine<br />

whether the adverse affects <strong>of</strong> habitat loss and fragmentation<br />

resulting from development are 'significant'. If they are<br />

determined to be significant, mitigation measures are required<br />

to reduce impacts below a level <strong>of</strong> significance. Unfortunately,<br />

because invertebrates typically receive little attention in the<br />

environmental review process, impacts to these species usually<br />

are undocumented.<br />

Management recommendations for the Hermes Copper<br />

include increased awareness <strong>of</strong> this species, particularly for<br />

those individuals and agencies involved in the environmental<br />

review process, and protection <strong>of</strong> existing colonies from habitat<br />

loss and fragmentation associated with urbanisation. The Hermes<br />

Copper is presently recognized by the United States Fish and<br />

Wildlife Service as a category 2 candidate species for listing as<br />

endangered or threatened; the Service recently received a<br />

121<br />

petition to list the species as threatened. Such a listing would<br />

increase significantly the protection afforded Hermes Copper.<br />

References<br />

BALMER, G. and PRATT, G. 1989 (1988). A survey <strong>of</strong> last instar larvae <strong>of</strong><br />

the <strong>Lycaenidae</strong> <strong>of</strong> California. J. Res. Lepid. 27: 1–70.<br />

BROWN, J.W. 1980. Hermes copper. Environment Southwest 491: 23. San<br />

Diego Natural History Museum.<br />

COMSTOCK, J.A. 1927. The <strong>Butterflies</strong> <strong>of</strong> California. Published by the<br />

author. 227 pp.<br />

COMSTOCK, J.A. and DAMMERS, CM. 1935. Notes on the early stages <strong>of</strong><br />

three butterflies and six moths from California. Bull. South. Calif. Acad.<br />

Sci. 34: 120–141.<br />

DOS PASSOS, C.F. 1964. A synonymic list <strong>of</strong> the Nearctic Rhopalocera.<br />

Mem. Lepid. Soc. 1: 1–145.<br />

EDWARDS, W.H. 1870. Description <strong>of</strong> a new species <strong>of</strong> Lepidoptera found<br />

within the United States. Trans. Amer. Entomol. Soc. 3: 21.<br />

EHRLICH, P.R. and EHRLICH, A.H. 1961. How to Know the <strong>Butterflies</strong>.<br />

Wm. C. Brown Co. Publ., Dubuque, Iowa. 262 pp.<br />

EMMEL, T.C. and EMMEL, J.F. 1973. <strong>Butterflies</strong> <strong>of</strong> Southern California.<br />

Nat. Hist. Mus. Los Angeles County, Sci. ser. 26: 1–148.<br />

GARTH, J. and TILDEN, J.W. 1986. California <strong>Butterflies</strong>. University <strong>of</strong><br />

California Press, Berkeley. 246 pp. + plates.<br />

HOWE, W.H. 1975. <strong>Butterflies</strong> <strong>of</strong> North America. Doubleday and Co., Inc.,<br />

Garden City, New York. 633pp.<br />

MILLER, L.D. and BROWN, F.M. 1979. A revision <strong>of</strong> the American coppers<br />

(Lepidoptera: <strong>Lycaenidae</strong>). Bull. Allyn Mus. 51: 22–23.<br />

ORSAK, L.J. 1977. <strong>Butterflies</strong> <strong>of</strong> Orange County. Univ. Calif. Irvine, Mus.<br />

Syst. Biol. Res. ser. 4. 349 pp.<br />

RINDGE, F.H. 1948. Contributions toward a knowledge <strong>of</strong> the insect fauna <strong>of</strong><br />

Lower California. No. 8. Lepidoptera: Rhopalocera. Proc. Calif. Acad.<br />

Sci. 24: 303.<br />

SCOTT, J.A. 1986. The butterflies <strong>of</strong> North America. Stanford University<br />

Press, Stanford, Calif. 583 pp.<br />

THORNE, F.T. 1963. The distribution <strong>of</strong> an endemic butterfly, Lycaena<br />

hermes. J. Res. Lepid. 2: 143–150.<br />

WRIGHT, W.G. 1906. <strong>Butterflies</strong> <strong>of</strong> the West Coast <strong>of</strong> the United States.<br />

Whitaker and Ray Co., San Francisco, Calif. 257 pp.<br />

WRIGHT, W.G. 1930. An annotated list <strong>of</strong> the butterflies <strong>of</strong> San Diego<br />

County, California. Trans. San Diego Soc. Nat. Hist. 6: 1–40.


Thome's Hairstreak, Mitoura thornei Brown<br />

John W. BROWN<br />

Entomology Department, San Diego Natural History Museum, P.O. Box 1390, San Diego, California 92112, U.S.A.<br />

Country: U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – possibly threatened.<br />

Thome's Hairstreak is a geographically isolated and<br />

ecologically distinct taxon that is restricted to a single mountain<br />

in southwestern San Diego County, California. Owing to its<br />

highly restricted distribution and the potential threats <strong>of</strong> habitat<br />

loss and degradation, Thorne's Hairstreak is recognized as a<br />

'category 2 candidate' for listing as endangered or threatened<br />

by the United States Fish and Wildlife Service.<br />

Taxonomy and Description: Thorne's Hairstreak was<br />

described by Brown (1983) as Mitoura thornei. Although most<br />

authors have treated it as a distinct species (e.g. Brown 1983;<br />

Garth and Tilden 1986; Ferris 1989; Ballmer and Pratt 1989),<br />

Shields (1984) suggests that it is a subspecies <strong>of</strong> Mitoura loki<br />

(Skinner), and Scott (1986) suggests that it is a subspecies <strong>of</strong><br />

Mitoura grynea Huebner. Most likely, M. thornei is part <strong>of</strong> a<br />

'superspecies' complex in which the degree <strong>of</strong> morphological<br />

divergence and genetic isolation among taxa do not conform<br />

well with our fixed system <strong>of</strong> binomial (or trinomial)<br />

nomenclature. Regardless <strong>of</strong> taxonomic opinion, Thome's<br />

Hairstreak is ecologically distinct and geographically isolated<br />

from its nearest congeners.<br />

The nearctic genus Mitoura Scudder frequently is considered<br />

a subgenus <strong>of</strong> Callophrys Billberg. Hence, Thome's Hairstreak<br />

occasionally is referred to as Callophrys(Mitoura) thornei.<br />

Distribution: Thome's Hairstreak is restricted to Otay Mountain<br />

(= San Ysidro Mountains) in the southwestern portion <strong>of</strong> San<br />

Diego County, California. On Otay Mountain it is confined to<br />

places where the larval foodplant, Tecate cypress (Cupressus<br />

forbesi; Cupressaceae), grows. Although a significant stand <strong>of</strong><br />

Tecate cypress occurs to the north in Coal Canyon, Orange<br />

County, California, and small populations are found to the<br />

south in northwestern Baja California, Mexico, Thorne's<br />

Hairstreak has not been documented from any <strong>of</strong> these localities<br />

(e.g. Orsak 1977; Brown 1983).<br />

Population Size: Otay mountain undoubtedly supports an<br />

extensive, nearly contiguous population (or set <strong>of</strong> populations)<br />

122<br />

<strong>of</strong> Thome's Hairstreak, although no quantitative data are<br />

available. Most <strong>of</strong> the Tecate cypress on the mountain has not<br />

been subject to encroachment by development or other human<br />

activities that result in loss <strong>of</strong> habitat, although chaparral fires<br />

frequently reduce or eliminate stands <strong>of</strong> the trees.<br />

Habitat and Ecology: Thorne's Hairstreak occurs only in<br />

southern interior cypress forest (Holland 1986) or where this<br />

habitat blends into other habitat types. The larval foodplant is<br />

Tecate cypress (Cupressus forbesi), a closed-cone conifer that<br />

occurs on mesic slopes and drainages in chaparral, and with<br />

which the adults are intimately associated. Thorne's Hairstreak<br />

is at least double brooded, with adults flying in late February<br />

through March, and again in June. Capture records indicate that<br />

the second brood may be only a partial one, as is the case with<br />

the closely related Mitoura loki. The emergence <strong>of</strong> laboratory<br />

reared individuals in August suggests the presence <strong>of</strong> a third<br />

brood in the fall, but this is yet to be documented in the field.<br />

The early stages <strong>of</strong> M. thornei closely resemble those<br />

described for M. siva (Edwards) (Coolidge 1924), M. loki<br />

(Comstock and Dammers 1932a), and M. nelsoni (Boisduval)<br />

(Comstock and Dammers 1932b), (see Ballmer and Pratt 1989).<br />

The eggs are echinoid and light green, and are laid singly on the<br />

new growth <strong>of</strong> the host plant. The egg stage lasts 7 to 14 days.<br />

Newly hatched larvae initially bore into the young stems <strong>of</strong> the<br />

host but later become external feeders. The larvae closely<br />

resemble the terminal twigs upon which they feed. Complete<br />

larval development, from hatching to pupation, requires 26–35<br />

days under laboratory conditions (Brown 1983). Pupation<br />

generally occurs in the duff or debris at the base <strong>of</strong> the host trees.<br />

No parasitoids or predators are recorded.<br />

Threats: Fire is an integral element in the natural history <strong>of</strong><br />

Tecate cypress as it is the major factor that initiates cone<br />

opening and subsequent seed dispersal (Zedler 1977). Chaparral<br />

fires undoubtedly have caused fluctuations in the populations<br />

<strong>of</strong> both the cypress and the associated butterfly. Zedler (1977)<br />

indicates that Tecate cypress requires approximately 25 years<br />

to reach reproductive maturity. Hence, an increase in the<br />

incidence <strong>of</strong> fire (i.e. a frequency <strong>of</strong> less than every 25 years)<br />

could severely affect the host trees. In recent years, chaparral


fires have been common in the Otay Mountain area, usually as<br />

a result <strong>of</strong> carelessness by people. Chaparral fires probably<br />

represent the greatest threat to the cypress and its associated<br />

insect fauna, including Thorne's Hairstreak.<br />

<strong>Conservation</strong>: Currently, Thome's Hairstreak receives minimal<br />

protection under the California Environmental Quality Act.<br />

Under this legislation, impacts to sensitive plants and animals<br />

and sensitive habitat types that result from development<br />

activities, must be assessed to determine whether the adverse<br />

affects <strong>of</strong> habitat loss and fragmentation are 'significant'. If<br />

they are determined to be significant, mitigation measures are<br />

required to reduce impacts below a level <strong>of</strong> significance.<br />

Unfortunately, because invertebrates typically receive little<br />

attention in the environmental review process, these impacts<br />

typically are undocumented.<br />

Management recommendations for Thorne's Hairstreak<br />

include increased awareness <strong>of</strong> this species, particularly to<br />

those individuals and agencies involved in the environmental<br />

review process, and increased fire control and fire management<br />

on Otay Mountain. Much <strong>of</strong> Otay Mountain is under the<br />

ownership <strong>of</strong> the United States Bureau <strong>of</strong> Land Management.<br />

Consequently, much <strong>of</strong> this land is likely to remain an open<br />

space. Thome's Hairstreak presently is recognized by the<br />

United States Fish and Wildlife Service as a category 2 candidate<br />

species for listing as endangered or threatened; the Service<br />

recently received a petition to list the species as threatened.<br />

Such a listing would increase significantly the protection<br />

afforded this species.<br />

123<br />

References<br />

BALLMER, G. and PRATT, G. 1989 (1988). A survey <strong>of</strong> last instar larvae <strong>of</strong><br />

the <strong>Lycaenidae</strong> <strong>of</strong> California. J. Res. Lepid. 27: 1–70.<br />

BROWN, J.W. 1983. A new species <strong>of</strong> Mitoura Scudder from southern<br />

California.J. Res. Lepid. 21: 245–254.<br />

COMSTOCK, J.A. and DAMMERS, CM. 1932a. Metamorphosis <strong>of</strong> five<br />

California diurnals (Lepidoptera). Bull. South Calif. Acad. Sci. 13:33–45.<br />

COMSTOCK, J.A. and DAMMERS, CM. 1932b. Metamorphosis <strong>of</strong> six<br />

California Lepidoptera. Bull. South Calif. Acad. Sci. 13: 88–100.<br />

COOLIDGE, K.R. 1924. The life history <strong>of</strong> Mitoura loki Skinner (Lepid.:<br />

<strong>Lycaenidae</strong>). Entomol. News 35: 199–204.<br />

FERRIS, C. 1989. Supplement to the catalogue/checklist <strong>of</strong> the butterflies <strong>of</strong><br />

America north <strong>of</strong> Mexico. Mem. Lepid. Soc. 3: 1–103.<br />

GARTH, J. and TILDEN, J.W. 1986. California <strong>Butterflies</strong>. University <strong>of</strong><br />

California Press, Berkeley. 246 pp.<br />

HOLLAND, R. 1986. Preliminary descriptions <strong>of</strong> the terrestrial natural<br />

communities <strong>of</strong> California. State <strong>of</strong> California, Department <strong>of</strong> Fish and<br />

Game. 156 pp.<br />

ORSAK, L.J. 1977. <strong>Butterflies</strong> <strong>of</strong> Orange County. Univ. Calif. Irvine, Mus.<br />

Syst. Biol. Res. ser. 4. 349 pp.<br />

SCOTT, J.A. 1986. The <strong>Butterflies</strong> <strong>of</strong> North America. Stanford University<br />

Press, Stanford, Calif. 583 pp.<br />

SHIELDS, O. 1984. Comments on recent papers regarding western<br />

Cupressaceae-feeding Callophrys (Mitoura). Utahensis 4: 51–56.<br />

ZEDLER, P.H. 1977. Life history attributes <strong>of</strong> plants and fire cycles; a case<br />

study in chaparral dominated by Cupressus forbesi. pp. 451–458. In:<br />

Mooney, H.A. and L.E. Conrad (tech. coords.), Proceedings <strong>of</strong> the<br />

Symposium on the Environmental Consequences <strong>of</strong> Fire and Fuel<br />

Management on Mediterranean Ecosystems. Palo Alto, Calif.


Sweadner's Hairstreak, Mitoura gryneus sweadneri (Chermock)<br />

Country: U.S.A.<br />

Thomas C. EMMEL<br />

Department <strong>of</strong> Zoology, University <strong>of</strong> Florida, Gainesville, Florida 32611, U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – rare, threatened.<br />

This local butterfly was described from the city <strong>of</strong> St.<br />

Augustine on the northeastern coast <strong>of</strong> Florida, a site which<br />

remains the centre <strong>of</strong> this rarity's distribution in the state.<br />

Virtually everywhere, it is threatened by reduced available<br />

habitat due to land clearing for housing and other development.<br />

In late 1986, after learning <strong>of</strong> the trend towards irreversible<br />

habitat loss for the Sweadner's Hairstreak, the mayor <strong>of</strong> St.<br />

Augustine led a drive to enact a city ordinance to protect the<br />

butterfly and its only known native foodplant. St. Augustine<br />

thus became the second city in the United States to protect a<br />

threatened butterfly site by law. (Pacific Grove on the central<br />

California coast protects the overwintering colonies <strong>of</strong> the<br />

common Monarch butterfly, Danaus plexippus.)<br />

Taxonomy and Description: The butterfly was described in<br />

1944 by noted lepidopterist Frank H. Chermock. Chermock<br />

named it as a distinct species, Mitoura sweadneri. A.B. Klots<br />

(1951) listed it as a subspecies <strong>of</strong> M. gryneus in his popular field<br />

guide, and it has been referred to as a subspecies ever since. In<br />

1993, Emmel, Baggett and Fee will publish a paper elevating<br />

the taxon to full specific status again, based on life history<br />

characteristics, hybrid crossing studies, and genitalic differences.<br />

Distribution: Sweadner's Hairstreak is considered very local<br />

and usually very rare. It has been found in several sites in the<br />

present city limits <strong>of</strong> Jacksonville, south through St. Augustine<br />

to New Smyrna Beach on the east coast <strong>of</strong> Florida, and several<br />

colonies are now known on the Gulf coast <strong>of</strong> Florida around<br />

Crystal River north to Cedar Key. Each colony is quite separated<br />

geographically from the others.<br />

Population Size and Status: All colonies are very isolated and<br />

generally quite small in number (from 12 to 50 adults normally<br />

being present). The adults <strong>of</strong> both sexes appear to occupy a<br />

small home range, generally on a single large cedar tree or on<br />

several closely adjacent cedars. However, the males are strongly<br />

territorial and chase other adults that enter their defended areas.<br />

124<br />

Thus, a maximum <strong>of</strong> three or four adults may be found spaced<br />

around a single tree.<br />

Habitat and Ecology: The butterfly lives in sandy coastal<br />

habitat occupied by its only known native foodplant, the Southern<br />

Red Cedar (Juniperus silicicola). These coniferous trees support<br />

both the immature stages and the adults, although the adults<br />

depend on locally growing wildflowers near the base <strong>of</strong> the<br />

trees for nectar.<br />

The egg <strong>of</strong> Sweadner's Hairstreak is pale green, and the<br />

young larva is superbly coloured to match the plant and scalelike<br />

leaves <strong>of</strong> the cedar twig. The mature larva is flat and sluglike<br />

in shape, and has a deep green ground colour with pale<br />

green diagonal stripes high on the sides. The pupa is dark brown<br />

and serves as the hibernating stage for the species during the<br />

short north Florida winter. The species has three annual broods<br />

(spring, summer, fall).<br />

For unknown reasons, the butterfly seems to be rather<br />

tightly adapted to coastal climates and is rarely found very far<br />

inland, despite the much wider inland distribution <strong>of</strong> Southern<br />

Red Cedars.<br />

Threats: The greatest threat to the species is continued<br />

development <strong>of</strong> the coastal habitat for housing resulting from<br />

urban expansion and the desire for recreational beach homes.<br />

Other threats include road construction, dump clearing, and<br />

related land disturbances which have destroyed many <strong>of</strong> the<br />

formerly available habitats for the species. Finally, even if the<br />

red cedar trees are left in the area by the government ordinance<br />

requirements, the multiple-brooded adults may not survive if<br />

the proper wildflowers are not left in close association with the<br />

surviving trees. Without nectar, the adults die within a day or<br />

two, prior to reproducing.<br />

<strong>Conservation</strong>: The conservation needs <strong>of</strong> Sweadner's<br />

Hairstreak lie not only in legal protection <strong>of</strong> its native foodplant<br />

but also in protecting strips <strong>of</strong> native vegetation associated with<br />

those Southern Red Cedar trees. The butterfly can survive in a<br />

remarkably small area (even 20m 2 ), as long as nectar sources<br />

are left in association with the red cedar trees. Popular support


in the cities <strong>of</strong> Jacksonville, St. Augustine, and St. Augustine<br />

Beach has led to the passing <strong>of</strong> several civic ordinances since<br />

1986 by those urban areas to preserve the butterfly and its<br />

foodplant. Additionally, the cities are reducing pesticide spraying<br />

for mosquito control in butterfly colony areas. There is<br />

considerable hope now that the loss <strong>of</strong> developmentally attractive<br />

coastal areas having Southern Red Cedar and Sweadner's<br />

Hairstreak colonies may be arrested (Emmel 1987).<br />

125<br />

References<br />

EMMEL, T.C. 1987. Delicate balance: Sweadner's Hairstreak butterfly<br />

(Mitoura gryneus sweadneri). Florida Wildlife 41 (2): 39.<br />

KLOTS, A.B. 1951. A Field Guide to the <strong>Butterflies</strong>. Houghton Mifflin Co.,<br />

Boston, Massachusetts. 349 pp.


Bartram's Hairstreak, Strymon acis bartrami (Corns tock &<br />

Huntington)<br />

Country: U.S.A.<br />

Thomas C. EMMEL and Marc C. Minno<br />

Department <strong>of</strong> Zoology, University <strong>of</strong> Florida, Gainesville, Florida 32611, U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status–rare; indeterminate<br />

(Red List).<br />

Bartram's hairstreak is classified as a threatened species<br />

because <strong>of</strong> its restricted distribution, low abundance, and recent<br />

loss <strong>of</strong> habitat.<br />

Taxonomy and Description: This subspecies is restricted to<br />

Florida and was described in 1943 by Comstock & Huntington.<br />

It is characterised by the heaviest white markings <strong>of</strong> all the<br />

subspecies <strong>of</strong> Strymon acis. The species ranges from Florida<br />

throughout the West Indies to Dominica (Riley 1975). Typical<br />

Strymon acis from Antigua and Dominica form the largest race.<br />

5. a. petioni Comstock & Huntington occurs on Hispaniola and<br />

has gray undersides. 5. a. mars Fabricius from the Virgin<br />

Islands, St. Kitts, and Puerto Rico has brown on the underside<br />

and a large orange area. S. a. gossei Comstock & Huntington<br />

from Jamaica and the Cayman Island resembles the last<br />

subspecies in the narrowness <strong>of</strong> the white bands, but the<br />

submarginal zone is broader. The Bahamas subspecies armouri<br />

Clench has narrow white bands and a narrow submarginal<br />

marking area. The Cuban subspecies casasi Comstock &<br />

Huntington is very similar to bartrami in Florida, with heavy<br />

white markings. The scientific and common name <strong>of</strong> Bartram's<br />

Hairstreak for the Florida population honours the memory <strong>of</strong><br />

William Bartram, an early Florida naturalist whose journeys<br />

through the state first brought wide recognition <strong>of</strong> its unique<br />

natural history.<br />

Distribution: This attractive little hairstreak is found only in<br />

southern Dade County, including the Everglades National<br />

Park, and on Big Pine Key in the Florida Keys. Hurricane<br />

Andrew may have severely impacted the mainland population<br />

<strong>of</strong> this butterfly on 24 August 1992, as it moved directly through<br />

the known colonies <strong>of</strong> this butterfly when crossing southern<br />

Dade County.<br />

Population Size: Thirty field surveys, conducted by Hennessey<br />

and Habeck (1991) between 23 May and 16 December 1988,<br />

found an average <strong>of</strong> 0.5 adults per hectare in pineland habitats<br />

126<br />

<strong>of</strong> the Everglades National Park and 0.3,1.0 and 2.7 individuals<br />

per hectare at three locations on Big Pine Key. The status <strong>of</strong><br />

mainland populations following Hurricane Andrew is not yet<br />

known at the time <strong>of</strong> writing (December 1992).<br />

Habitat and Ecology: Bartram's Hairstreak is a small grayish<br />

butterfly with two pairs <strong>of</strong> delicate tails on the hindwings. The<br />

undersides <strong>of</strong> the hindwings have a highly distinctive pattern <strong>of</strong><br />

white spots and lines, plus a red eyespot at the base <strong>of</strong> the tail.<br />

It occurs in open tropical pinelands that have an abundance <strong>of</strong><br />

the larval hostplant. Females lay their eggs singly on the<br />

flowers <strong>of</strong> Woolly Croton, Croton Hnearis. The larvae feed on<br />

the flowers and leaves <strong>of</strong> the host. Adults frequently perch on<br />

Woolly Crotons and visit nearby flowers for nectar. Several<br />

generations are produced each year.<br />

Threats: The primary threats to this species are housing<br />

developments and agricultural clearing in the heavily crowded<br />

southern Dade County area. However, the natural disaster<br />

presented by the tremendously destructive Hurricane Andrew<br />

winds, on 24 August 1992, may have destroyed or severely<br />

affected all the remaining habitat areas on the mainland. On<br />

both Big Pine Key in the Florida Keys and in southern Dade<br />

County, open tropical pinelands are also threatened periodically<br />

by fire.<br />

<strong>Conservation</strong>: Additional surveys are badly needed now to<br />

monitor the abundance and distribution <strong>of</strong> Bartram's Hairstreak<br />

in both southern Dade County and the Florida Keys. Ecological<br />

studies should be conducted to identify the species' habitat<br />

requirements. Prescribed burning <strong>of</strong> pinelands may be necessary<br />

to maintain habitat for this species (Hennessey and Habeck<br />

1991), but land managers should take care not to burn large<br />

tracts entirely, lest populations <strong>of</strong> Bartram's Hairstreak and<br />

other rare butterflies be destroyed by the fire. Careful<br />

management planning is needed for the remaining habitat in the<br />

Keys and in the Everglades National Park. The species is an<br />

attractive butterfly and could be used in publicity campaigns<br />

advocating preservation <strong>of</strong> these increasingly rare tropical<br />

pineland habitats in Florida (Minno and Emmel 1993).


References<br />

HENNESSEY, M.K. and HABECK, D.H. 1991. Effects <strong>of</strong> mosquito adulticides<br />

on populations <strong>of</strong> non-target terrestrial arthropods in the Florida Keys.<br />

Unpublished final report. U.S. Fish and Wildlife Service and University<br />

127<br />

<strong>of</strong> Florida Cooperative Wildlife Research Unit, Gainesville, Florida.<br />

MINNO, M.C. and EMMEL, T.C. 1993. <strong>Butterflies</strong> <strong>of</strong> the Florida Keys.<br />

Scientific Publishers, Gainesville, Florida. 168 pp.<br />

RILEY, N.D. 1975. A Field Guide to the <strong>Butterflies</strong> <strong>of</strong> the West Indies.<br />

Demeter Press, Inc., Boston, Massachusetts. 224 pp.


The Avalon Hairstreak, Strymon avalona (W.G. Wright)<br />

Thomas C. EMMEL and John F. EMMEL<br />

Department <strong>of</strong> Zoology, University <strong>of</strong> Florida, Gainesville, Florida 32611, U.S.A., and 26500 Rim Road, Hemet, California<br />

92544, U.S.A.<br />

Country: U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – insufficiently<br />

known (Wells et al. 1983, Red List).<br />

This hairstreak is one <strong>of</strong> the world's most restricted species,<br />

being found only on Santa Catalina Island just slightly more<br />

than 20 miles <strong>of</strong>f the coast from Los Angeles, southern California.<br />

The limited land available for development in southern<br />

California, including Santa Catalina, means that ultimately this<br />

species is threatened by development. Proximate dangers include<br />

overgrazing and other destruction by cattle and feral livestock<br />

such as goats.<br />

Possible hybridisation with the related S. melinus (which<br />

was recorded from Santa Catalina for the first time in 1978) has<br />

been suggested, but such claims are probably inaccurate (Gall<br />

1985; Gorelick 1987).<br />

Taxonomy and Description: This species was described by<br />

W.G. Wright in 1905 from the vicinity <strong>of</strong> Avalon on Catalina<br />

Island. It does not occur on any <strong>of</strong> the other Channel Islands,<br />

where a related species, the Gray Hairstreak (Strymon melinus<br />

Hü+bner) may be found. No geographic variation has been noted<br />

in Strymon avalona on Santa Catalina Island.<br />

Distribution: This species occurs only on Santa Catalina<br />

Island, <strong>of</strong>f the coast <strong>of</strong> southern California.<br />

Population Size: The Avalon Hairstreak is common in select<br />

localities such as the hills around Avalon. The butterfly is<br />

generally distributed in various localities from an elevational<br />

range <strong>of</strong> sea level to 65m. A number <strong>of</strong> colonies have been<br />

reported along the road to Renton Mine, Pebbly Beach, Jewfish<br />

Point, the Isthmus, and hills to the west <strong>of</strong> Avalon.<br />

Habitat and Ecology: The species prefers chaparral and grasscovered<br />

slopes at relatively low elevations. Captures have been<br />

reported in every month <strong>of</strong> the year, with the first brood primarily<br />

occurring from mid-February through April, a second brood in<br />

July and August, and a third brood flying from September to<br />

November. The adults frequently perch on bushes and grassy<br />

areas in chaparral, and visit the flowers <strong>of</strong> common sumac (Rhus<br />

alurind) and the giant buckwheat (Eriogonum giganteum Wats.).<br />

128<br />

Eggs are laid singly, usually in the terminal buds or on<br />

immature flowers <strong>of</strong> the foodplant, Lotus argophyllus (Gray)<br />

Greene var. ornithopus (Greene) Ottley, and Lotus scoparius<br />

(Nutt.) Ottley, in the Leguminosae. Mature larvae show<br />

considerable variation in ground colour, ranging from a pale<br />

apple green to pale pink. The entire body <strong>of</strong> the larva, except the<br />

cervical shield, is covered with short white fine hairs. The pupa<br />

is pale pinkish-brown or wood brown with markings <strong>of</strong> various<br />

shades <strong>of</strong> olive-green. Pupation occurs at the base <strong>of</strong> the<br />

foodplant, with the usual support <strong>of</strong> a delicate silk girdle<br />

(Emmel, Emmel and Mattoon in press).<br />

Threats: The species is not known to be endangered at the<br />

present time, although increased urbanisation or changes in<br />

vegetation cover due to overgrazing by domestic and feral<br />

livestock constitute potential threats every year. No special<br />

conservation measures have been taken on Santa Catalina<br />

Island for this butterfly.<br />

<strong>Conservation</strong>: The widespread continental species Strymon<br />

melinus has not yet become successfully established on Santa<br />

Catalina Island. Likewise, this rare endemic Strymon avalona<br />

has not established itself on any <strong>of</strong> the other Channel Islands or<br />

on the mainland <strong>of</strong> North America. The situation should be<br />

monitored yearly to better understand why Strymon melinus in<br />

particular, has not managed to establish itself yet on Santa<br />

Catalina Island where many <strong>of</strong> its potential foodplants are<br />

growing.<br />

References<br />

EMMEL, T.C., EMMEL, J.F. and MATTOON, S.O. In press. The <strong>Butterflies</strong><br />

<strong>of</strong> California. Stanford University Press, Stanford, California.<br />

GALL, L.F. 1985. Santa Catalina's endemic Lepidoptera. II. The Avalon<br />

Hairstreak, Strymon avalona, and its interaction with the recently<br />

introduced Gray Hairstreak, Strymon melinus (<strong>Lycaenidae</strong>). In: Menke,<br />

A.S. and Miller, D.R. (Eds). Proc. 1st Symposium on Entomology <strong>of</strong> the<br />

California Islands. Santa Barbara Museum <strong>of</strong> Natural History, pp. 95–104.<br />

GORELICK, G.A. 1987. Santa Catalina's endemic Lepidoptera. II. The<br />

Avalon Hairstreak, Strymon avalona (<strong>Lycaenidae</strong>): an ecological study.<br />

Atala 14 (1986): 1–12.<br />

WELLS, S.M., PYLE, R.M. and COLLINS, N.M. 1983. The <strong>IUCN</strong> Invertebrate<br />

Red Data Book. <strong>IUCN</strong>, Gland, Switzerland.


Country: U.S.A.<br />

The Atala Butterfly, Eumaeus atala florida (Röber)<br />

Thomas C. EMMEL and Marc C. MINNO<br />

Department <strong>of</strong> Zoology, University <strong>of</strong> Florida, Gainesville, Florida 32611, U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – out <strong>of</strong> danger;<br />

vulnerable (Red List).<br />

The Atala butterfly is one <strong>of</strong> the most strikingly coloured<br />

butterflies in Florida and it is desirable to collectors. Its<br />

conservation interest arises from the fact that the species was<br />

feared extirpated in the early 1970s after the few known<br />

colonies died out, but then made a spectacular recovery starting<br />

in 1979 from a single small colony rediscovered in the Miami<br />

area. The species has recovered much <strong>of</strong> its former range and is<br />

now even considered a pest <strong>of</strong> ornamental cycad plantings! It is<br />

assessed as relatively common in parts <strong>of</strong> southern Florida<br />

(Bowers and Larin 1989; Minno and Emmel 1993).<br />

This species was the subject <strong>of</strong> one <strong>of</strong> three full lycaenid<br />

entries in the <strong>IUCN</strong> Invertebrate Red Data Book (Wells et al.<br />

1983), and one <strong>of</strong> six butterflies discussed in the Invertebrate<br />

volume <strong>of</strong> Franz (1982). It gave its name 'Atala' to a journal<br />

published by the Xerces Society.<br />

Taxonomy and Description: The subspecies was described by<br />

Röber in 1926, on the basis <strong>of</strong> having more extensive bluegreen<br />

on the upper side and larger spots <strong>of</strong> that colour on the<br />

underside <strong>of</strong> the hindwings. The typical subspecies, Eumaeus<br />

atala atala Poey (1832), occurs in Cuba, on Andros Island, and<br />

on Great Abaco Island in the Bahamas (Riley 1975). Some<br />

lepidopterists have questioned if the Florida population is<br />

sufficiently distinct to be called a separate race (Clench 1977).<br />

Distribution: The Atala Butterfly is found in southern Florida<br />

from Broward County in the vicinity <strong>of</strong> Fort Lauderdale south<br />

to southern Dade County, and it is recorded historically from<br />

Elliott Key and Key Largo (the most recent record for the latter<br />

Key being 5 June 1960). Most <strong>of</strong> the populations in southern<br />

Dade County were probably severely impacted by Hurricane<br />

Andrew on 24 August 1992.<br />

Population Size: The Atala Butterfly was once abundant in the<br />

rimrock areas <strong>of</strong> the southern mainland <strong>of</strong> Florida, but largescale<br />

harvesting <strong>of</strong> the host plant, coontie (Zamia pumila,<br />

Cycadaceae) for starch in the late 1800s greatly reduced the<br />

129<br />

number <strong>of</strong> coontie plants. Urbanisation and development <strong>of</strong> the<br />

coastal habitat favoured by the Atala also had a large impact. By<br />

1965, the Atala had been reduced to a single known population<br />

in Hugh Taylor Birch State Park. After this colony died out, the<br />

Atala was feared extirpated from Florida. However, in the late<br />

1970s, another colony was found on Virginia Key in the Miami<br />

area. <strong>Conservation</strong>ists such as Roger Hamler <strong>of</strong> Dade County<br />

Parks set out potted coontie plants on which females laid eggs.<br />

Plants with eggs were then moved to other locations and new<br />

colonies were started. Rawson (1961) demonstrated the<br />

possibility <strong>of</strong> translocating E. atala by liberating adults in a new<br />

site.<br />

The Atala has made a spectacular recovery and is now found<br />

in urban and natural areas around Fort Lauderdale and Miami,<br />

and has been successfully introduced into the Everglades<br />

National Park. Some plant nurseries and botanical gardens<br />

currently consider the Atala a pest species, as the larvae are<br />

capable <strong>of</strong> defoliating lamia species used in landscaping. It is<br />

not known whether our current population is <strong>of</strong> original Florida<br />

stock or the result <strong>of</strong> a new introduction from the Bahamas or<br />

Cuba.<br />

The very few records <strong>of</strong> Atala from the Florida Keys include<br />

only Elliot Key and Key Largo (Schwartz 1888). Small (1913)<br />

listed coontie among the plants found in the Keys, but we have<br />

not encountered it on any <strong>of</strong> the islands. The early pioneers<br />

probably extirpated coontie (and thereby the butterfly) from the<br />

Keys, as it was a readily available source <strong>of</strong> starch. The only<br />

modern record <strong>of</strong> the Atala from the Keys is the single capture<br />

<strong>of</strong> a male on 5 June 1960 in the City <strong>of</strong> Key Largo. The status<br />

<strong>of</strong> the populations on the mainland was probably severely<br />

impacted by Hurricane Andrew on 24 August 1992, and<br />

subsequent surveys have not yet been done to ascertain the<br />

species' status on the mainland.<br />

Habitat and Ecology: The Atala is found in tropical pinelands<br />

and hardwood hammocks in close association with the larval<br />

foodplant, coontie (Zamia pumila, Cycadaceae). The Atala<br />

adults are the largest lycaenids in Florida and occur all year<br />

round.It is one <strong>of</strong> the most strikingly coloured butterflies on its<br />

ventral surfaces, with jet black wings, iridescent blue spots and<br />

a red patch on the underside <strong>of</strong> the hindwings. The upper wings


are black with iridescent green in males and iridescent blue in<br />

females, while the abdomen is bright red. The adults have a<br />

slow fluttering flight pattern. Males perch on the leaves <strong>of</strong><br />

shrubs and make circular flights around the perch site like other<br />

hairstreaks. Both sexes <strong>of</strong>ten visit flowers.<br />

The white eggs are laid in clusters on the young growth <strong>of</strong><br />

coontie. Larvae are bright red with a yellow spot, and are<br />

probably distasteful to birds or other predators. Pupae are<br />

brown with small dark spots and hang from the substrate by a<br />

silk girdle. Droplets <strong>of</strong> bitter tasting liquid are exuded over the<br />

cuticle <strong>of</strong> the pupae.<br />

Threats: The Atala is listed as a species <strong>of</strong> special concern in<br />

Florida because <strong>of</strong> its restricted distribution and cyclic<br />

fluctuations in abundance. This butterfly is currently known<br />

from more colonies on the mainland than were recorded before<br />

1965, but the Atala has lost habitat areas formerly occupied in<br />

the Florida Keys. Hurricane Andrew probably severely impacted<br />

populations <strong>of</strong> the Atala in southern Dade County.<br />

Continued threats occur in both wild and urban areas. The Atala<br />

occurs in tropical hardwood hammocks and pinelands in close<br />

association with the host plant. It also now uses some urban<br />

areas such as gardens and nurseries where the foodplant is<br />

grown as an ornamental. Thus this species is exposed to both<br />

habitat clearing and burning in the wild (although prescribed<br />

burning <strong>of</strong> pinelands may be necessary to maintain habitat).<br />

Spraying or other pest control measures in urban habitats pose<br />

another threat. For example, before Hurricane Andrew, Fairchild<br />

Botanical Garden in southern Dade County regularly used Bti<br />

sprays to control the infestation <strong>of</strong> the Atala Butterfly on its<br />

valuable cycad collection (from the 1980s through summer<br />

1992).<br />

130<br />

<strong>Conservation</strong>: The unexpected occurrence <strong>of</strong> wide habitat<br />

destruction by Hurricane Andrew in the summer <strong>of</strong> 1992<br />

necessitates new surveys to monitor the distribution and<br />

abundance <strong>of</strong> the Atala in Florida. Outside <strong>of</strong> the affected urban<br />

areas, prescribed burning <strong>of</strong> pinelands may be necessary to<br />

maintain habitat for Atalas in natural settings. The species is<br />

probably permanently lost from the Florida Keys, unless<br />

restoration plantings <strong>of</strong> the coontie host plants are done on Big<br />

Pine Key where some substantial tropical pinelands remain<br />

(there are no pinelands left on Elliot Key or Key Largo, its<br />

originally recorded habitat in the Florida Keys). Taxonomic<br />

studies should be conducted to determine if the current taxon<br />

present in south Florida is the same as that present before 1965,<br />

and to define the relationship <strong>of</strong> E. atala florida to E. atala atala<br />

in the Bahamas.<br />

References<br />

BOWERS, M.D. and LARIN, Z. 1989. Acquired chemical defense in the<br />

lycaenid butterfly, Eumaeus atala. J. Chem. Ecol. 15: 1133–1146.<br />

CLENCH, H.K. 1977. A list <strong>of</strong> the butterflies <strong>of</strong> Andros, Bahamas. Ann.<br />

Carnegie Mus. 46: 173–194.<br />

FRANZ, R. (Ed.) 1982. Rare and Endangered Biota <strong>of</strong> Florida. 6. Invertebrates.<br />

University Presses <strong>of</strong> Florida.<br />

MINNO, M.C. and EMMEL, T.C. 1993. <strong>Butterflies</strong> <strong>of</strong> the Florida Keys.<br />

Scientific Publishers, Gainesville, Florida. 168 pp.<br />

RAWSON, G.W. 1961. The recent rediscovery <strong>of</strong> Eumaeus atala (<strong>Lycaenidae</strong>)<br />

in southern Florida. J. Lepid. Soc. 15: 237–244.<br />

RILEY, N.D. 1975. A Field Guide to the <strong>Butterflies</strong> <strong>of</strong> the West Indies.<br />

Demeter Press, Inc., Boston, Massachusetts. 224 pp.<br />

SCHWARTZ, E.A. 1888. Notes on Eumaeus atala. Insect Life 1: 37–40.<br />

SMALL, J.K. 1913. Flora <strong>of</strong> the Florida Keys. Published by the author, New<br />

York. 162 pp.<br />

WELLS, S.M., PYLE, R.M. and COLLINS, N.M. 1983. The<strong>IUCN</strong>Invertebrate<br />

Red Data Book. <strong>IUCN</strong>, Gland.


Smith's Blue, Euphilotes enoptes smithi (Mattoni)<br />

Thomas C. EMMEL and John F. EMMEL<br />

Department <strong>of</strong> Zoology, University <strong>of</strong> Florida, Gainesville, Florida 32611, U.S.A., and 26500 Rim Road, Hemet,<br />

California 92544, U.S.A.<br />

Country: U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered (Red<br />

List).<br />

This butterfly is known only from the coastal fog belt <strong>of</strong><br />

Monterey County in California, where it inhabits the immediate<br />

coast <strong>of</strong> the Big Sur country. A member <strong>of</strong> a widely distributed<br />

western U.S. species, Euphilotes enoptes smithi is an endemic<br />

California subspecies whose coastal habitat has suffered a<br />

number <strong>of</strong> disturbances, including beach recreation and <strong>of</strong>froad<br />

vehicles. Montane habitats, in general, have suffered less<br />

(Arnold 1983a,b).<br />

Taxonomy and Description: The subspecies was described<br />

(Mattoni 1954) as different from the numerous other subspecies<br />

<strong>of</strong> Euphilotes enoptes in part because <strong>of</strong> the broad black<br />

marginal borders on the lustrous blue upper wings <strong>of</strong> the males.<br />

Females are brown above, with a band <strong>of</strong> red-orange marks<br />

across the hindwings. The overall distinguishing features from<br />

other subspecies are the light undersurface ground colour and<br />

prominent black markings with a faint black terminal line<br />

(Emmel, Emmel and Mattoon in press).<br />

The male is distinguished by the broad marginal border <strong>of</strong><br />

the hindwings. The ventral surface <strong>of</strong> both males and females<br />

has a faint terminal line and a light ground colour with large<br />

prominent spots.<br />

Distribution: This species is confined to coastal Monterey<br />

County from Big Sur and the mouth <strong>of</strong> the Salinas River<br />

southward to Del Rey Creek and an area several miles north <strong>of</strong><br />

the San Luis Obispo County line. It ranges from near sea level<br />

to approximately 65m elevation. The type locality is at Burns<br />

Creek, State Highway 1, in Monterey County.<br />

Population Size: Populations are small in number and probably<br />

have never been particularly large because <strong>of</strong> the relatively<br />

restricted habitat.<br />

Habitat and Ecology: The Smith's Blue butterfly occurs on<br />

cliffs, steep slopes, and road cuts along the immediate coast,<br />

131<br />

within the northern coastal scrub plant community. It also is<br />

found extensively on coastal and inland sand dunes, and<br />

occasionally on serpentine grassland.<br />

The adults emerge between mid-June and early September,<br />

corresponding with the blooming <strong>of</strong> the buckwheat plants on<br />

which they feed, rest, sun, and mate. While each adult lives for<br />

only about one week, individual emergences are scattered over<br />

the extended flight period. Females deposit eggs singly in<br />

buckwheat flowers. Larvae hatch 4–8 days later and go through<br />

five instars before pupating in flowerheads or in the litter and<br />

sand at the base <strong>of</strong> the plant. Pupae hang in place from September<br />

until adults emerge the following year. The verified host plant<br />

for the larvae is Eriogonum parvifolium Sm. in Rees. (Pratt and<br />

Ballmer 1989). Larvae also feed on E. latifolium, and differences<br />

in plant phenology at different sites may represent a potential<br />

isolating mechanism for populations.<br />

Threats: The primary threat to the conservation and recovery<br />

<strong>of</strong> the Smith's Blue butterfly is the wide variety <strong>of</strong> man-made<br />

disturbances <strong>of</strong> the coastal habitat. Dunes are widely threatened<br />

by beach recreation, <strong>of</strong>f-road vehicles, housing developments,<br />

and road construction. Additionally, non-native plants such as<br />

iceplant and Holland dunegrass invade the dunes and displace<br />

native buckwheat. At Fort Ord, the sand dunes have been<br />

damaged by military vehicles and infantry exercises. At the<br />

Seaside-Marina dune system and the Del Monte Forest, their<br />

habitat has been destroyed by sand mining. More than half the<br />

dune habitat present at the turn <strong>of</strong> the century had been destroyed<br />

by about 1980 (Powell 1981).<br />

<strong>Conservation</strong>: A conservation plan for Smith's Blue designed<br />

by Arnold (1983b) was one <strong>of</strong> the first detailed prescriptions for<br />

a North American lycaenid. Prevention <strong>of</strong> further habitat loss or<br />

change was the prime need, and Arnold' s plan aimed to maintain<br />

known populations <strong>of</strong> E. e. smithi by coordinating habitat<br />

preservation, rehabilitation and management.<br />

Five categories <strong>of</strong> action were proposed: (1) preservation<br />

and protection <strong>of</strong> existing habitats; (2) implementation <strong>of</strong> shortterm<br />

and long-term management; (3) development <strong>of</strong> monitoring<br />

programmes to census selected populations annually to assess


the effects <strong>of</strong> management efforts; (4) promotion <strong>of</strong> public<br />

awareness <strong>of</strong> the butterfly and its habitat, and (5) enforcement<br />

<strong>of</strong> laws and regulations to protect the butterfly.<br />

This species was listed as Endangered on June 1, 1976. In<br />

1977, the U.S. Army established a butterfly preserve at Fort<br />

Ord, and some <strong>of</strong> the non-native plants have been removed<br />

there as well as native plants re-established. For remnant sand<br />

dune habitats at Sand City in Monterey County, the city agreed<br />

to complete a conservation plan before proceeding with any<br />

work in the dune areas that have been zoned for housing<br />

development.<br />

132<br />

References<br />

ARNOLD, R.A. 1983a. Ecological studies on six endangered butterflies<br />

(Lepidoptera: <strong>Lycaenidae</strong>): Island biogeography, patch dynamics, and<br />

the design <strong>of</strong> habitat preserves. Univ. Calif. Publns Entomol. 99: 1–161.<br />

ARNOLD, R.A. 1983b. <strong>Conservation</strong> and management <strong>of</strong> the endangered<br />

Smith's Blue Butterfly, Euphilotes enoptes smithi (Lepidoptera:<br />

<strong>Lycaenidae</strong>).J. Res. tepid. 22: 135–153.<br />

EMMEL, T.C., EMMEL, J.F. and MATTOON, S.O. In press. The <strong>Butterflies</strong><br />

<strong>of</strong> California. Stanford University Press, Stanford, California.<br />

MATTONI, R.H.T. 1954. Notes on the genus Philotes. 1. Description <strong>of</strong> three<br />

new subspecies and a synoptic list. Bull. South. Calif. Acad. Sci. 53:<br />

157–165.<br />

POWELL, J.A. 1981. Endangered habitats for insects: California coastal sand<br />

dunes. Atala 6: 41–55.<br />

PRATT, G.P. and BALLMER, G.R. 1989. A survey <strong>of</strong> the last instar larvae <strong>of</strong><br />

the <strong>Lycaenidae</strong> (Lepidoptera) <strong>of</strong> California. J. Res. Lepid. 27: 1–81.


The El Segundo Blue, Euphilotes bernardino allyni (Shields)<br />

Country: U.S.A.(California).<br />

Status and <strong>Conservation</strong> Interest: Status – endangered (Red<br />

List).<br />

This subspecies is a sand obligate ecotype found only on the<br />

El Segundo sand dune ecosystem <strong>of</strong> the coastal plain <strong>of</strong> western<br />

Los Angeles. Its remaining three discrete colonies are surrounded<br />

by a dense urban area that remains both a fast growing and fast<br />

denaturing area. As a federally listed endangered species, the El<br />

Segundo Blue (ESB) is significant in conferring an umbrella <strong>of</strong><br />

protection on at least ten other species <strong>of</strong> plant and animals that<br />

are restricted to this sand dune system and to more than 20<br />

species restricted to the coastal sand dunes <strong>of</strong> southern California<br />

and northern Baja California.<br />

Taxonomy and Description: The subspecies was described by<br />

Shields (1975) from the El Segundo sand dunes <strong>of</strong> Los Angeles<br />

county just prior to its listing among the first group <strong>of</strong> butterflies<br />

to be legally recognised as endangered by the Endangered<br />

Species Act <strong>of</strong> 1973. The taxon was originally classified as a<br />

subspecies <strong>of</strong> E. battoides, but both Shields (Shields 1975,<br />

Reveal and Shields 1988) and Mattoni (1989) independently<br />

diagnosed allyni as more logically belonging to the bernardino<br />

group <strong>of</strong> four related but clearly separated subspecies which<br />

they recognised as a distinct species.<br />

Distribution: The ESB was historically restricted to the El<br />

Segundo sand dunes that covered about 1200ha. Today three<br />

discrete colonies exist on fragments that still maintain some<br />

characteristics <strong>of</strong> the natural community.<br />

Population Size: The largest population is on property <strong>of</strong> the<br />

Los Angeles International Airport (LAX), where 80ha were<br />

recently set aside by the city <strong>of</strong> Los Angeles as a biological<br />

preserve, in part to satisfy biotic requirements <strong>of</strong> the butterfly.<br />

When serious study started in 1984, total population size was<br />

about 500. Following restoration efforts the standing population<br />

in 1990 was about 4000 (Mattoni 1988, 1990a, b, 1992). The<br />

second largest colony is on a 0.6ha lot on the Chevron Oil<br />

refinery property. The population prior to study was about 2000<br />

and has since decreased to under 500 (Arnold 1983,1986). A<br />

R.H.T. MATTONI<br />

9620 Heather Road, Beverly Hills, California 90210, U.S.A.<br />

133<br />

small colony <strong>of</strong> a few hundred butterflies persists on a 0.2ha<br />

isolated dune fragment at Redondo Beach. The latter was<br />

discovered in 1984.<br />

Habitat and Ecology: The species is sand obligate and adapted<br />

to a single foodplant, the coastal buckwheat, Eriogonum<br />

parvifolium, (Polygonaceae). It has one generation per year,<br />

adults appearing from June through mid-August. Excepting the<br />

fossorial diapausing pupal stage, the entire life cycle <strong>of</strong> the<br />

butterfly is associated with flowerheads <strong>of</strong> the buckwheat<br />

including egg deposition, larval growth, nectaring, mating and<br />

dying (Mattoni 1991). The larvae are ant associated, but the<br />

relationship is facultative and the only ant now noted in<br />

association is the exotic Argentine ant, Iridomyrmex humilis.<br />

Adult butterflies were highly sedentary at the tiny Chevron site<br />

(Arnold 1983), but nothing is known <strong>of</strong> movements at LAX<br />

beyond observed concentrations near dense patches <strong>of</strong> foodplant.<br />

Threats: Pratt (1987) first recognised that the major threat to<br />

the ESB at LAX was the presence <strong>of</strong> a dense common buckwheat<br />

stand which had been planted during a poorly conceived<br />

restoration effort 20 years earlier. This earlier flowering species<br />

provided sustenance to a guild <strong>of</strong> non-diapausing Lepidoptera<br />

that later migrated to the flowering coastal buckwheat.<br />

Overwhelming numbers, cannibalism, and shared parasites<br />

devastated the ESB. Removal <strong>of</strong> part <strong>of</strong> the exotic plant biomass<br />

was believed a major cause <strong>of</strong> the recent resurgence <strong>of</strong> ESB<br />

populations (Mattoni 1991).<br />

The ESB at the Chevron site, while serving as a linchpin in<br />

the oil company's advertising campaign to show its concern for<br />

nature, precipitously declined (Arnold 1986). The cause was<br />

probably an intensive mark-release programme to assess<br />

population size, coupled with extreme trampling across a small<br />

area. The Redondo Beach site is completely untended, but is not<br />

suitable for building development and is thus not threatened by<br />

developers.<br />

<strong>Conservation</strong>: Listing the ESB under the federal Endangered<br />

Species Act provides one <strong>of</strong> the greatest success stories <strong>of</strong> that<br />

legislation. For the principal colony at LAX, the 1976 listing<br />

happened just in time to halt a plan to develop almost the entire


120ha <strong>of</strong> dunes as a golf course. Following a protracted planning<br />

and political process, the preserve area increased in size. Under<br />

leadership <strong>of</strong> the local Councilwoman, the Mayor and Airport<br />

Commission agreed to set aside 80ha <strong>of</strong> the highest quality land<br />

as a permanent preserve with the proviso that the golf course on<br />

the remaining 40ha be designed with all rough areas re-planted<br />

habitat.<br />

In the meantime, the Airport Commission provided $180,000<br />

to begin restoration <strong>of</strong> the dunes ecosystem. Work on the dunes<br />

is now in its third year.<br />

By contrast, the Chevron company has emphasised creating<br />

conditions to maximise survival <strong>of</strong> the ESB, at the expense <strong>of</strong><br />

the ecosystem. The main effort has been directed towards<br />

creating a butterfly garden; because the site is so small, this<br />

approach has some validity. On the other hand long-term<br />

stability would be better insured with a community mimicking<br />

that <strong>of</strong> the historic system, and not a near monoculture <strong>of</strong> the<br />

foodplant with more glamorous advertising potential. Oppewall<br />

(1975) documents efforts by local collectors who successfully<br />

convinced Chevron to set aside the area as a preserve. For this,<br />

the company is to be commended.<br />

134<br />

References<br />

ARNOLD, R.A. 1983. Ecological studies <strong>of</strong> six endangered butterflies<br />

(Lepidoptera, <strong>Lycaenidae</strong>): Island biogeography, patch dynamics, and<br />

the design <strong>of</strong> habitat preserves. Univ. Calif. Publns Entomol. 99. 161pp.<br />

ARNOLD, R.A. 1986. Private and government funded conservation programs<br />

for endangered insects in California. Natural Areas Journal 5: 28–39.<br />

MATTONI, R.H.T. 1988. Captive propagation <strong>of</strong> California endangered<br />

butterflies. Report to Calif. Dept. Fish and Game. Contract C-1456.<br />

MATTONI, R.H.T. 1989. The Euphilotes battoides complex: Recognition <strong>of</strong><br />

a species and description <strong>of</strong> a new subspecies. J. Res. Lepid. 27:173–185.<br />

MATTONI, R.H.T. 1990a. Habitat evaluation and species diversity on the<br />

LAX El Segundo sand dunes. Rept. to the LAX board <strong>of</strong> airport<br />

commissioners.<br />

MATTONI, R.H.T. 1990b. Unnatural acts: succession on the El Segundo<br />

sand dunes in California. In: Hughes, H.G. and T.M. Bonnickson (Eds.)<br />

Proc. first SER Conference, Berkeley, CA 1989. Soc. Ecol. Restoration,<br />

Madison WI 53711.<br />

MATTONI, R.H.T. 1992. The endangered El Segundo blue butterfly. J. Res.<br />

Lepid. 29: 277–304 (1990).<br />

OPPEWALL, J.C. 1975. The saving <strong>of</strong> the El Segundo Blue. Atala 3: 25–8.<br />

PRATT, G. 1987. Competition as a controlling factor <strong>of</strong> Euphilotes battoides<br />

allyni larval abundance. Atala 15: 1–9.<br />

SHIELDS, O. 1975. Studies on North American Philotes IV. Taxonomic and<br />

biological notes and new subspecies. Bull. Allyn Museum No. 28. 36pp.<br />

SHIELDS, O. and REVEAL, J. 1988. Sequential evolution <strong>of</strong> Euphilotes<br />

(<strong>Lycaenidae</strong>, Scolitantidini) on their plant host Eriogonum (Polygonaceae,<br />

Eriogonoideae). J. Linn. Soc. 33: 51–91.


The Palos Verdes Blue, Glaucopsyche lygdamus palosverdesensis<br />

Perkins and Emmel<br />

Country: U.S.A. (California)<br />

Status and <strong>Conservation</strong> Interest: Status – probably extinct<br />

(Red List).<br />

The subspecies, now certainly extinct, was a coastal bluff<br />

ecotype found only on the southern half <strong>of</strong> the Palos Verdes<br />

peninsula in southern Los Angeles county. The species has high<br />

conservation value, nonetheless, as it has not been <strong>of</strong>ficially<br />

delisted under the theory that extinction is not certain. Hence all<br />

building projects in the species distribution area – and there are<br />

many – must recognise habitat value. At least one major<br />

development is currently stalled waiting approval <strong>of</strong> a plan to<br />

protect foodplants. The great benefit <strong>of</strong> the situation is provision<br />

<strong>of</strong> protection to other endangered species, now unlisted,<br />

occurring in the habitat.<br />

Taxonomy and Description: The subspecies was described by<br />

Perkins and Emmel (1977) from Los Angeles county just prior<br />

to its listing among the second group <strong>of</strong> butterflies to be legally<br />

recognised as endangered by the Endangered Species Act. The<br />

taxon was diagnosed as a subspecies distinct from, Glaucopsyche<br />

lygdamus australis, the southern blue, by exclusive use <strong>of</strong> the<br />

milk vetch Astragalus leucopsis, very fast flight, and several<br />

wing characteristics.<br />

Distribution. By the time <strong>of</strong> its discovery by Perkins in the<br />

early 1970s, the Palos Verdes Blue (PVB) was already restricted<br />

to a few fragments retaining some natural characteristics. In<br />

1977 at least nine discrete colonies existed. The last known<br />

occurrence was in 1983.<br />

Population Size: The largest populations known during the<br />

brief time span the PVB was extant were at Atala Vista Terrace<br />

(type locality) and in the scrub extending from Palos Verdes<br />

Drive east to Friendship Park. The former locality was built<br />

over in 1978. Population sizes were never estimated and by the<br />

early 1980s numbers were extremely low, probably less than<br />

100 adults among all the remaining fragments at that time<br />

(Arnold 1985). In spring <strong>of</strong> 1982 at Hesse Park, I counted six<br />

adults on the best day, with some 20 plants. Each plant had at<br />

R.H.T. MATTONI<br />

9620 Heather Road, Beverly Hills, California 90210, U.S.A.<br />

135<br />

least 100 eggs, and one plant over 500. Foodplant availability<br />

was limiting, probably due to spring disking for fire control.<br />

Habitat and Ecology: The PVB was a coastal, sage-associated<br />

ecotype, restricted in foodplant use to the milk vetch. The vetch<br />

is restricted to the fog belt across southern exposures at elevations<br />

between 100 and 300 metres. The historical area probably<br />

occupied by the PVB was no more than 4000ha. The flora <strong>of</strong> the<br />

northwest slopes <strong>of</strong> the peninsula included the low shrub<br />

legume, Lotus scoparious, foodplant <strong>of</strong> the sister subspecies,<br />

G. lygdamus australis. Whether australis was parapatric to<br />

palosverdesensis is unknown.<br />

The butterfly was single brooded with adult flight in February<br />

and March. Eggs were usually laid on flowerheads <strong>of</strong> the<br />

foodplant, but when foodplant numbers were diminished just<br />

prior to extinction, eggs were laid over the entire plant. Larvae<br />

usually fed on seed within developing seedpods <strong>of</strong> the vetch and<br />

were ant tended in the last two instars. The final known<br />

generation, observed at Hesse Park, had larvae feeding on<br />

leaves as well since the flowerheads and seeds were exhausted.<br />

Three other Lycaenid butterflies were associated with the<br />

flowerhead/seedpod guild: Strymon melinus; Leptotes marina;<br />

and Everes amyntula. The first two are polyphagous, have<br />

many alternative foodplants, and are widespread species. The<br />

latter, with the PVB is a monophage restricted to the vetch. It<br />

has been extirpated from the Palos Verdes peninsula, although<br />

the last specimens were sighted in 1986 (Jess Morton, Tony<br />

Leigh, pers. comm.).<br />

Threats Leading to Extinction. Arnold (1986) reported the<br />

decline <strong>of</strong> the species and speculated that it had become extinct.<br />

Intensive search has been conducted by several local<br />

lepidopterists every year since 1983 without success. The<br />

proximate cause <strong>of</strong> extinction was denaturing <strong>of</strong> the land by<br />

development and fire suppression tactics. The historic population<br />

must have been continuous over the 4000ha coastal scrub<br />

habitat. With intensive development from 1950 the habitat was<br />

fragmented, although one 500ha section remains. Clearing<br />

practice so degraded this that the construction at Hesse Park in<br />

1982, performed by the city <strong>of</strong> Rancho Palos Verdes in violation


<strong>of</strong> the federal Endangered Species Act, destroyed the last<br />

remaining colony. The city was subsequently sued by the<br />

federal government, but this legal action was dismissed under<br />

the theory that the city could not be held liable.<br />

<strong>Conservation</strong>. Ironically, a simple captive breeding method<br />

was developed in 1983 using the southern blue as a surrogate<br />

(Mattoni 1988). Had the technique been available the previous<br />

year, it would have been possible to have saved the species by<br />

captive breeding for later re-introduction into the habitat.<br />

Unfortunately the foodplant stands were continually being<br />

cleared for fire prevention and few plants now remain. It has<br />

been suggested (Mattoni unpublished) that an effort be<br />

implemented to release numbers <strong>of</strong> the southern blue, preadapted<br />

to feeding on vetch by captive mass rearing, into sites to be<br />

heavily restocked with vetch. This action would again grace the<br />

136<br />

area with a butterfly that may evolve characteristics <strong>of</strong> its<br />

extinct relative. In addition to restoring natural biodiversity the<br />

plan would provide a demonstration <strong>of</strong> adaptive processes in a<br />

restored environment.<br />

References<br />

ARNOLD, R.A. 1985. Palos Verdes blue butterfly management plan. U.S.F.<br />

and W.S. informal report. 37pp.<br />

ARNOLD, R.A. 1986. Decline <strong>of</strong> the endangered Palos Verdes blue butterfly<br />

in California. Biol. Conserv. 40: 203–217.<br />

MATTONI, R.H.T. 1988.1988 report to California Department <strong>of</strong> Fish and<br />

Game, Contract C-1456.<br />

PERKINS, E.M. and EMMEL, J.F. 1977. A new subspecies <strong>of</strong> Glaucopsyche<br />

lygdamus from California (Lepidoptera: <strong>Lycaenidae</strong>). Proc. Ent. Soc.<br />

Wash. 79: 468–471.


The Xerces Blue, Glaucopsyche xerces (Boisduval)<br />

Thomas C. EMMEL and John F. EMMEL<br />

Department <strong>of</strong> Zoology, University <strong>of</strong> Florida, Gainesville, Florida 32611, U.S.A., and 26500 Rim Road, Hemet, California<br />

92544, U.S.A.<br />

Country: U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – extinct (Red List).<br />

The Xerces Blue, a former resident <strong>of</strong> the sand dunes in San<br />

Francisco, was the first butterfly species in North America to<br />

become extinct through human interference. (Two satyrines in<br />

the genus Cercyonis were the first subspecies <strong>of</strong> wider ranging<br />

U.S. species to disappear in historic times.) Today, an<br />

international organization devoted to the conservation <strong>of</strong><br />

invertebrates, The Xerces Society, is named for this diminutive<br />

creature, a butterfly which has achieved fame far beyond its size<br />

and former restricted geographic distribution. From a scientific<br />

viewpoint, it was one <strong>of</strong> the most variable butterflies known and<br />

it would have made a unique tool for research on the effects <strong>of</strong><br />

population size and geographic distribution on infra-specific<br />

variation. But for its untimely extinction, G. xerces would have<br />

contributed greatly to human knowledge in ecology and<br />

evolutionary biology.<br />

Taxonomy and Description: The species was first described<br />

by Boisduval (1852) from one male and two females taken<br />

within the area now occupied by the city <strong>of</strong> San Francisco, in<br />

San Francisco County, California. These specimens were<br />

collected by Pierre Joseph Michel Lorquin, a French gold<br />

seeker and naturalist who arrived in San Francisco in late 1849<br />

or early 1850 and apparently made his first shipment <strong>of</strong> specimens<br />

to Boisduval in the fall <strong>of</strong> 1851. Thus he had only two seasons<br />

at the most to collect the first set <strong>of</strong> specimens that he sent to<br />

Boisduval. While the species was extraordinarily variable, and<br />

several forms were named (originally thought to be separate<br />

species), all are presently referred to the nominotypical<br />

subspecies (Emmel, Emmel and Mattoon in press) and no<br />

geographic subspecies are recognized.<br />

The closest living relative <strong>of</strong> the Xerces may well be a<br />

newly recognized subspecies <strong>of</strong> Glaucopsyche lygdamus<br />

residing on Santa Rosa Island, <strong>of</strong>f the coast <strong>of</strong> southern California<br />

(Emmel and Emmel in press). The extraordinary phenotypic<br />

resemblance <strong>of</strong> the wing maculation <strong>of</strong> this new subspecies to<br />

that <strong>of</strong> G. xerces suggests a close phylogenetic relationship.<br />

The genitalia <strong>of</strong> G. xerces are very similar to those <strong>of</strong> G.<br />

lygdamus, and on the basis <strong>of</strong> male genitalia alone, the Xerces<br />

137<br />

Blue would have been assigned subspecific status under<br />

lygdamus. However, Downey and Lange (1956) pointed out<br />

considerable differences between these species in larval stages,<br />

adult wing maculation, and ecology. Additionally, the two<br />

species were once sympatric in San Francisco and hybrids were<br />

never detected.<br />

Distribution: All recorded specimens are from San Francisco,<br />

and include the area presently covered by the city. The<br />

distribution on the San Francisco Peninsula ranged from near<br />

Twin Peaks to North Beach, and from the Presidio on the Bay<br />

southward to Lake Merced. Lone Mountain, formerly an isolated,<br />

sandy hill, was the classical locality for Xerces in the early<br />

1900s.<br />

Population Size: The only known colonies were in the San<br />

Francisco Bay area. No estimates <strong>of</strong> population size have come<br />

down to us in the early literature, but by comparison with the<br />

closely related G. lygdamus colonies in the area, we can guess<br />

that the typical size was several hundred individuals in a<br />

colony. By the year 1919, the only known population remaining<br />

was flying in a limited area west <strong>of</strong> the Marine Hospital, at the<br />

Presidio, San Francisco. The same area was still inhabited on<br />

March 23, 1941. The butterflies were then limited to a small<br />

area 21m wide by 46m long, in which Lotus scoparius was<br />

found. The last known specimens <strong>of</strong> Xerces were collected at<br />

the Presidio during May 1941 (Downey and Lange 1956), and<br />

many later visits to the area in search <strong>of</strong> the butterfly were<br />

fruitless.<br />

Habitat and Ecology: The habitat <strong>of</strong> the PVB was sandy areas<br />

where a prostrate dune ecotype <strong>of</strong> the perennial legume, Lotus<br />

scoparius (Nutt.) Ottley, occurred. The butterfly typically<br />

occurred around patches <strong>of</strong> Lotus that grew in the partial shade<br />

<strong>of</strong> the Monterey Cypress (Cupressus macrocarpa Hartw.) in<br />

well-drained, sandy soil. Lupinus arboreus Sims, which also<br />

served the Xerces females as a hostplant for oviposition and<br />

larval growth, was widely distributed among the Lotus, and had<br />

a wider range than that <strong>of</strong> the Lotus and the butterflies. In 1956,<br />

the same association <strong>of</strong> plants at the last colony site remained<br />

essentially unchanged from 1941, and was also observed by the


authors in 1963–67 in essentially unchanged condition. To the<br />

north, similar sandy areas with these Lotus and Lupinus species<br />

occur in Marin County, but the Monterey Cypress does not<br />

occur there. It appears that the expansion <strong>of</strong> the city <strong>of</strong> San<br />

Francisco into the natural sand dune habitat formerly available<br />

to Glaucopsyche xerces (and the other insects <strong>of</strong> the region) had<br />

destroyed too much <strong>of</strong> the habitat for the butterfly to continue<br />

to support a sustainable population beyond 1941.<br />

The adults <strong>of</strong> Xerces flew in a single brood from about<br />

March 10 to April 15. However, specimens were taken from as<br />

early as late February to as late as early June. The flight, mating<br />

behaviour, and oviposition behaviour <strong>of</strong> Xerces were apparently<br />

similar to that <strong>of</strong> the other Glaucopsyche species <strong>of</strong> California.<br />

The female laid eggs singly in the small depression at the base<br />

<strong>of</strong> the leaflet on Lupinus arboreus, or laid eggs on new growth<br />

near the tips <strong>of</strong> the leaflets on Lotus scoparius. In 1939, it was<br />

observed that as many as nine eggs were laid per plant, and that<br />

the Lotus was slightly favoured by the females, with an average<br />

<strong>of</strong> 3–4 eggs per plant compared to 2 eggs per plant for the<br />

Lupinus.<br />

The butterflies were associated with ants in their natural<br />

habitat although the ant associates were never identified. In the<br />

laboratory, larvae were fed substitute foodplants (Lupinus<br />

micranthus Dougl. and Astragalus menziessii Gray), and the<br />

captive larvae were raised without any ants so the species was<br />

not completely dependent on ants for successful maturation and<br />

pupation. The pale green larvae had long hairs on each side <strong>of</strong><br />

the dorsal surface, and the whole body was covered with a<br />

whitish pile. The general colouration and pattern were variable,<br />

a trait found in other nearctic blue species today. The pupal<br />

colouration was highly variable. Diapause was in the pupal<br />

stage, with the length <strong>of</strong> that stage averaging from 10 to 11<br />

months.<br />

138<br />

Threats: The decline <strong>of</strong> this fascinating butterfly and its final<br />

extinction in 1941 appears to be attributable solely to the<br />

expansion <strong>of</strong> the City <strong>of</strong> San Francisco in the preceding 60–80<br />

years. This resulted in the removal <strong>of</strong> the native vegetation in<br />

the dunes and reduction <strong>of</strong> the remaining habitat areas to an<br />

unsustainable size. In such small populations, a minor change<br />

in climate or other environmental factor could prove devastating,<br />

especially to a species that flew in a single annual brood and had<br />

no options to react, little time or space, and a small genetic pool<br />

<strong>of</strong> variability. Collecting might have been detrimental also as<br />

the populations declined.<br />

While a live Xerces Blue will never be seen again on Earth,<br />

the species has served as a remarkably effective symbol for the<br />

fragility <strong>of</strong> nature among American citizens and conservationists<br />

around the world. Scientific analysis <strong>of</strong> 344 historical specimens<br />

(Downey and Lange 1956) <strong>of</strong>fer a hint as to the remarkable<br />

variation in this species and the annual changes <strong>of</strong> the frequency<br />

<strong>of</strong> pattern forms in the population.<br />

References<br />

BOISDUVAL, J.B.A.D. de. 1852. Lepidoptères de la Californie. Ann. Soc.<br />

Entomol. France (Series 2) 10: 275–324.<br />

DOWNEY, J.C. and LANGE, JR., W.H. 1956. Analysis <strong>of</strong> variation in a<br />

recently extinct polymorphic Lycaenid butterfly, Glaucopsyche xerces<br />

(Bdv.), with notes on its biology and taxonomy. Bull. South. Calif. Acad.<br />

Sci. 55 (3): 153–207.<br />

EMMEL, T.C. and EMMEL, J.F. In press. A new xerces-like subspecies <strong>of</strong><br />

Glaucopsyche (<strong>Lycaenidae</strong>) from Santa Rosa Island, California.<br />

Systematics <strong>of</strong> Western North American <strong>Butterflies</strong>. Scientific Publishers,<br />

Gainesville, Florida.<br />

EMMEL, T.C, EMMEL, J.F. and MATTOON, S.O. In press. The <strong>Butterflies</strong><br />

<strong>of</strong> California. Stanford University Press, Stanford, California.


The Mission Blue, Plebejus icarioides missionensis Hovanitz<br />

J. HALL CUSHMAN<br />

Center for <strong>Conservation</strong> <strong>Biology</strong>, Department <strong>of</strong> Biological Sciences, Stanford University, Stanford, CA 94305, U.S.A.<br />

Country: U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered (Red<br />

List).<br />

In 1976, the Mission Blue was <strong>of</strong>ficially listed as an<br />

endangered species by the U.S. Fish and Wildlife Service. The<br />

subspecies is also listed as endangered with the California<br />

Department <strong>of</strong> Fish and Game and almost all <strong>of</strong> its habitat<br />

protected as park land.<br />

Taxonomy and Description: The Mission Blue is one <strong>of</strong> 12<br />

recognized subspecies <strong>of</strong> the highly variable Plebejus icarioides<br />

and was first described by Hovanitz (1937).<br />

Distribution: While Plebejus icarioides as a whole is patchily<br />

distributed throughout western North America, the present-day<br />

distribution <strong>of</strong> the Mission Blue subspecies is restricted to four<br />

known areas on the northern tip <strong>of</strong> San Francisco Peninsula.<br />

Although it is impossible to document the historical distribution<br />

<strong>of</strong> this subspecies, the Mission Blue almost certainly occurred<br />

throughout much <strong>of</strong> the coastal scrub habitat <strong>of</strong> the northern<br />

peninsula (Reid and Murphy 1986).<br />

Population Size: The largest known Mission Blue population<br />

is on San Bruno Mountain (San Mateo Co.) and is estimated to<br />

consist <strong>of</strong> 18,000 adults. A substantial population is also found<br />

at Fort Baker (Marin Co.), although a precise estimate <strong>of</strong> its size<br />

is not available. Significantly smaller populations are found on<br />

Twin Peaks (San Francisco Co.) and the Skyline Ridges (San<br />

Mateo Co.), with estimates <strong>of</strong> 500 and 2000 adults, respectively<br />

(Reid and Murphy 1986).<br />

Habitat and ecology: The Mission Blue has one generation per<br />

year, and adults fly from mid-April to mid-June. Adults live for<br />

up to one week, remain close to the larval host plant, and feed<br />

on nectar from a wide variety <strong>of</strong> plants, including Eriogonum<br />

(Polygonaceae) and numerous composites. Females oviposit<br />

on three lupine species, Lupinus albifrons, L. formosus, and L.<br />

varicolor, which are found in areas <strong>of</strong> recent disturbance and<br />

attain highest densities in grasslands on thin, rocky soils (Arnold<br />

1983; Reid and Murphy 1986).<br />

139<br />

Females lay eggs on the leaves, stems, flowers, and seed<br />

pods <strong>of</strong> Lupinus hosts. The eggs are usually deposited singly<br />

and hatch in 4–10 days. After about three weeks, the secondinstar<br />

larvae enter an obligatory diapause from which they<br />

emerge the following spring. These post-diapause larvae<br />

subsequently complete their development to adult in four to<br />

five weeks (Arnold 1983; Reid and Murphy 1986; Newcomer<br />

1911).<br />

The cryptically coloured larvae may exhibit significant agespecific<br />

differences in feeding behaviour. Downey (1962)<br />

noted that older larvae are most commonly found at the base <strong>of</strong><br />

food plants while the younger larvae are substantially higher up<br />

on the host. Because <strong>of</strong> this, he postulated that the majority <strong>of</strong><br />

mature larvae in this species are nocturnal feeders.<br />

Juvenile stages <strong>of</strong> the Mission Blue are attacked by a variety<br />

<strong>of</strong> natural enemies. Arnold (1983) reported that 'Approximately<br />

35% <strong>of</strong> field collected eggs were parasitised by an unidentified<br />

encyrtid wasp.' In addition, Newcomer (1911) and Downey<br />

(1962) both reported that the larval stages <strong>of</strong> P. icarioides were<br />

parasitized by various braconid and tachinid species.<br />

As with many other lycaenid species, Mission Blue larvae<br />

are tended by ants. These associations have not received much<br />

attention (although see Downey 1962), and are thus poorly<br />

understood. However, the larvae are known to possess abdominal<br />

nectary glands which become active in the third or fourth instar<br />

and attract ants. Downey (1962) found that the butterfly could<br />

be reared successfully in the laboratory without ants, but makes<br />

no statements about how <strong>of</strong>ten older larvae are ant-tended in the<br />

field. Clearly, the importance <strong>of</strong> ants to the development and<br />

survival <strong>of</strong> Mission Blue larvae is very much an open question<br />

and requires further study.<br />

Threats: Agricultural and urban expansion have resulted in the<br />

progressive loss <strong>of</strong> native grassland habitat and reduced exchange<br />

among existing Mission Blue populations. These major threats<br />

have been significantly reduced due to the habitat conservation<br />

plan mentioned in the following section. However, a remaining<br />

threat is the invasion <strong>of</strong> grassland habitats by non-native plant<br />

species. While lupine, the Mission Blue host plant, is relatively<br />

resistant to invasions <strong>of</strong> non-native grasses, it is quite susceptible<br />

to the invasion <strong>of</strong> woody species which create too much shade.


The major culprits in this regard are gorse (Ulex europeaus),<br />

blue gum (Eucalyptus globulus) and broom (Cytisus spp.), all<br />

<strong>of</strong> which are on the increase in grassland areas such as those on<br />

San Bruno Mountain.<br />

<strong>Conservation</strong>: Concern about this subspecies, as well as the<br />

San Bruno Elfin (Callophrys mossi bayensis) and Callippe<br />

Silverspot (Speyeria callippe), was instrumental in leading to<br />

the formation <strong>of</strong> the U.S.A.'s first habitat conservation plan in<br />

1983. The plan created the San Bruno Mountain County Park,<br />

an extensive area encompassing the largest existing butterfly<br />

population. Considerable effort has been made to annually<br />

monitor the size <strong>of</strong> adult populations and initiate management<br />

activities such as the control <strong>of</strong> invasive species (see Reid and<br />

Murphy 1986, and Bean et al. 1991 for details).<br />

140<br />

References<br />

ARNOLD, R.A. 1983. Ecological studies <strong>of</strong> six endangered butterflies<br />

(Lepidoptera, <strong>Lycaenidae</strong>): island biogeography, patch dynamics, and<br />

design <strong>of</strong> habitat preserves. Univ. Calif. Publns Entomol. 99: 1–161.<br />

BEAN, M., FITZGERALD, S. and O'CONNOR, M. 1991. The San Bruno<br />

habitat conservation plan. In: Reconciling conflicts under the endangered<br />

species act World Wildlife Fund, Washington, D.C. pp. 52–65.<br />

DOWNEY, J.C. 1962. Myrmecophily in Plebejus (Icaricia) icarioides<br />

(Lepidoptera: <strong>Lycaenidae</strong>). Entomol. News 73: 57–66.<br />

HOVANITZ, W. 1937. Concerning the Plebejus icarioides Rassenkreis<br />

(Lepidoptera: <strong>Lycaenidae</strong>). Pan-Pacific Entomol. 13: 184–189.<br />

NEWCOMER, E.J. 1911. The life histories <strong>of</strong> two lycaenid butterflies. Can.<br />

Ent. 43: 83–88.<br />

REID, T.S. and MURPHY, D.D. 1986. The endangered Mission Blue butterfly,<br />

Plebejus icarioides missionensis. In: B.A. Wilcox, P.F. Brussard and<br />

B.G. Marcot (Eds). The management <strong>of</strong> viable populations: theory,<br />

applications, and case studies. Center for <strong>Conservation</strong> <strong>Biology</strong>, Stanford,<br />

California, pp. 147–167.


The San Bruno Elfin, Incisalia mossii bayensis (Brown)<br />

Stuart B. WEISS<br />

Center for <strong>Conservation</strong> <strong>Biology</strong>, Department <strong>of</strong> Biological Sciences, Stanford University, Stanford, CA 94305, U.S.A.<br />

Country: U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered (Red<br />

List).<br />

This subspecies is endemic to the northern San Francisco<br />

Peninsula, California. It was one <strong>of</strong> the first butterflies protected<br />

under the U.S. Endangered Species Act in 1976. As a local<br />

endemic in a highly urbanised region, threats to the San Bruno<br />

Elfin include land development and invasive introduced plant<br />

species.<br />

Taxonomy and Description: The subspecies was discovered<br />

rather late, in 1962 (MacNeill 1963), and described by Brown<br />

(1969a). It was originally described as Callophrysfotis bayensis,<br />

but was later recognized to be in the species C. mossii (Edwards)<br />

(now genus Incisalia, which occurs from Vancouver Island<br />

along the coast range to near Los Angeles). Populations to the<br />

north in Marin County are recognized as a different subspecies,<br />

those to the south in the Santa Lucia Range are ssp. doudor<strong>of</strong>fi.<br />

Distribution: The subspecies is restricted to three distinct areas<br />

on the northern San Francisco peninsula: Montara Mountain;<br />

Milagra Ridge; and San Bruno Mountain. Each <strong>of</strong> these localities<br />

supports an array <strong>of</strong> highly local demographic units tied together<br />

by occasional adult migration. Populations probably once existed<br />

within San Francisco at Twin Peaks and Mount Davidson, but<br />

have disappeared with urbanisation (Emmel and Ferris 1972).<br />

Population Size: The San Bruno Elfin was never common,<br />

because <strong>of</strong> specialized habitat requirements (see below). The<br />

butterfly exists in local discrete populations <strong>of</strong> ten to several<br />

hundred adults at higher altitudes. A thousand or more adults<br />

may exist in about 15 total subpopulations on San Bruno<br />

Mountain in a good year. Montara Mountain supports about 10<br />

local populations, and Milagra Ridge supports about four.<br />

Virtually all <strong>of</strong> the existing habitat is now protected as parkland,<br />

and numerous populations have been qualitatively monitored<br />

since 1982. Colonies noted by Arnold (1984) occupied small<br />

areas (0.15–8.0ha) on steep, north-facing slopes.<br />

141<br />

Habitat and ecology: The distribution <strong>of</strong> the butterfly closely<br />

follows the narrow, fragmented distribution <strong>of</strong> its larval<br />

hostplant, stonecrop, Sedum spathulifolium (Brown 1969b).<br />

This succulent plant grows in abundance only on thin-soiled or<br />

rocky north-facing slopes within the coastal fog belt. Sedum<br />

occurs in both short-statured coastal scrub and grassland<br />

vegetation types, and is most common near the summits <strong>of</strong><br />

coastal mountains and around rocky outcrops on lower slopes.<br />

Sedum readily invades roadcuts and old quarry faces provided<br />

the aspect is correct. Local populations <strong>of</strong> the Elfin correspond<br />

closely to these patches <strong>of</strong> the larval hostplant, which range<br />

from a hundred square meters to several hectares in extent.<br />

San Bruno Elfin adults fly from February into April, during<br />

the latter part <strong>of</strong> the rainy season in northern California, but<br />

before the onset <strong>of</strong> persistent summer fog. Adults usually<br />

appear after the first extended warm sunny period <strong>of</strong> the season,<br />

as early as the first week in February, or as late as April. The<br />

window <strong>of</strong> sunny, calm conditions during the flight season is<br />

highly variable from year to year, and adults run the risk <strong>of</strong><br />

being grounded by inclement weather for weeks on end.<br />

Populations were greatly reduced during and after near record<br />

rainfall in 1983, but appear to be less affected by recent drought<br />

conditions (San Bruno Mountain <strong>Conservation</strong> Plan Monitoring<br />

Reports, 1982–1991).<br />

Habitat topography may be limiting for Elfin populations in<br />

certain cases. Because <strong>of</strong> low winter sun angles, the steepest<br />

habitat areas may be in deep shade for much <strong>of</strong> the day, limiting<br />

access by adults (Weiss and Murphy 1991). While natural<br />

contours are rarely shaded all day even in February, roadcuts<br />

and quarry faces may face severe shading limitations. Steep<br />

(>40°) north-facing slopes provide minimal solar exposure<br />

even in March, and are rarely occupied by Elfin. Equally steep<br />

northeast-facing slopes receive direct morning light when winds<br />

are calm, and provide excellent Elfin habitat. Elfin activity on<br />

steep northwest-facing slopes, however, may be limited by<br />

strong afternoon winds.<br />

Adults are highly sedentary, typically moving less than 100<br />

metres (Arnold 1983), with a maximum recorded movement <strong>of</strong><br />

about 800m (Arnold 1984). Males perch on rocks and vegetation,<br />

and dart out at passing insects; females spend much <strong>of</strong> their time<br />

crawling among the foodplants (Emmel and Ferris 1972; Arnold


1983). Both sexes visit flowers (especially Lomatium<br />

utriculatum, but also other early blooming coastal species)<br />

when plants are available, but a number <strong>of</strong> local Elfin populations<br />

on Montara Mountain exist where nectar resources are<br />

astonishingly sparse.<br />

The butterfly is univoltine. Females oviposit on Sedum<br />

rosettes. Early instars feed on the fleshy leaves, but third and<br />

fourth instars feed on the flowers when they appear. Third and<br />

fourth instars are easily observed basking and feeding on<br />

Sedum flowerheads, and exhibit a continuous colour<br />

polymorphism ranging from deep red (the colour <strong>of</strong> Sedum<br />

foliage) through orange to bright yellow (the colours <strong>of</strong> Sedum<br />

flowers); larvae may change colour morph over a few days<br />

(Orsak and Whitman 1986). Larvae may be tended by ants <strong>of</strong> up<br />

to nine different species, but the relationship appears facultative<br />

(Arnold 1983). However, most larvae in the field are observed<br />

without ants (Emmel and Ferris 1972, S.B. Weiss pers. obs.).<br />

Larvae are parasitized by a tachinid, Aplomya theclarum that<br />

emerges from the fourth instar Elfin larvae. Parasitisation rates<br />

<strong>of</strong> reared larvae and collected pupae are high, <strong>of</strong> the order <strong>of</strong><br />

50–80% (Arnold 1983). Given the high densities <strong>of</strong> larvae<br />

observed (one or more per square meter), such high mortality<br />

appears necessary to produce the typically low density <strong>of</strong> adults<br />

the following year. The fourth instar caterpillar pupates in the<br />

duff immediately below the hostplants, and diapause lasts<br />

through the summer, fall and early winter.<br />

Threats: Because the vast majority <strong>of</strong> San Bruno Elfin<br />

populations are on public land (including San Bruno Mountain<br />

County Park, Golden Gate National Recreation Area, and<br />

McNee Ranch State Park) further opportunities for habitat<br />

conversion are limited. A proposed six lane road would skirt a<br />

population on Montara Mountain, but that construction has<br />

been challenged on a host <strong>of</strong> environmental and development<br />

issues other than the San Bruno Elfin. Continued expansion <strong>of</strong><br />

a quarry on San Bruno Mountain could destroy some habitat on<br />

that property; however, the quarry is scheduled to be shut down<br />

within a decade. The prohibitions <strong>of</strong> the Endangered Species<br />

Act and park rules provide a strong deterrent against<br />

overcollecting by both amateurs and scientists. Wildfires may<br />

threaten in more heavily vegetated areas, but the thin vegetation<br />

on rocky outcrops is relatively safe from fires. Vegetation<br />

succession also appears unimportant, given the thin soils and<br />

windswept conditions <strong>of</strong> the habitat. Invasive introduced species<br />

such as gorse (Ulex europeaus), brooms (Cytisus spp.), pampas<br />

grass, ice plant (Mesembryanthemum spp.), and blue gum<br />

(Eucalyptus globulus) are encroaching on some local<br />

populations.<br />

<strong>Conservation</strong>: Since most <strong>of</strong> the habitat is already protected,<br />

conservation prescriptions for the San Bruno Mountain fall<br />

under the jurisdiction <strong>of</strong> the San Bruno Mountain Habitat<br />

142<br />

<strong>Conservation</strong> Plan, 1982 (Bean et al. 1991), which includes<br />

yearly monitoring <strong>of</strong> adult numbers and management activities.<br />

Elements <strong>of</strong> a Recovery Plan (Arnold 1984) included: (1)<br />

protection <strong>of</strong> essential habitat (which was designated on each<br />

site), through a range <strong>of</strong> strategies including cooperative<br />

agreements, easements and others; (2) prevention <strong>of</strong> further<br />

habitat degradation, and habitat enhancement where possible,<br />

through minimising toxin use, removal <strong>of</strong> weeds, control <strong>of</strong> <strong>of</strong>froad<br />

vehicles and revegetating with native flora; (3) development<br />

and implementation <strong>of</strong> management plans for extant colonies<br />

by utilising annual surveys and fostering autecological studies;<br />

(4) re-establishment <strong>of</strong> the species in restored sites within its<br />

historical range; (5) increase in public awareness; and (6)<br />

enforcement and evaluation <strong>of</strong> protective laws and regulations<br />

at all levels.<br />

Control <strong>of</strong> invasive species is currently under way at Milagra<br />

Ridge and San Bruno Mountain. Large areas <strong>of</strong> Sedum along<br />

Wolf Ridge in Marin County (just outside the historical range<br />

<strong>of</strong> the Elfin) are presently unoccupied by the butterfly, raising<br />

the possibility <strong>of</strong> an introduction attempt there. Revegetating<br />

abandoned quarry faces on San Bruno Mountain would provide<br />

a great opportunity to increase the habitat <strong>of</strong> the San Bruno<br />

Elfin, because Sedum will rapidly invade bare rock surfaces.<br />

The quarry configuration would provide large areas <strong>of</strong><br />

appropriate solar exposure (especially northeast-facing slopes<br />

and flat benches) if restoration were attempted (Weiss and<br />

Murphy 1991).<br />

References<br />

ARNOLD, R.A. 1983. Ecological studies <strong>of</strong> six endangered butterflies<br />

(Lepidoptera, <strong>Lycaenidae</strong>): island biogeography, patch dynamics, and<br />

design <strong>of</strong> habitat preserves. Univ. Calif. Publns Entomol. 99: 1–161.<br />

ARNOLD, R.A. 1984. (Main author <strong>of</strong>) U.S. Fish and Wildlife Service. 1984.<br />

Recovery plan for the San Bruno Elfin and Mission Blue butterflies. U.S.<br />

Fish and Wildlife Service, Portland, Oregon.<br />

BEAN, M., FITZGERALD, S. and O'CONNOR, M. 1991. The San Bruno<br />

habitat conservation plan. In: Reconciling Conflicts under the Endangered<br />

Species Act. World Wildlife Fund, pp. 52–65.<br />

BROWN, R.M. 1969a. A new subspecies <strong>of</strong> Callophrys fotis from the San<br />

Francisco Bay Area. J. Lepid. Soc. 23: 95–96.<br />

BROWN, R.M. 1969b. Larva and habitat <strong>of</strong> Callophrys fotis bayensis.J. Res.<br />

Lepid. 8: 49–50.<br />

EMMEL, J.F. and FERRIS, CD. 1972. The biology <strong>of</strong> Callophrys fotis<br />

bayensis. J. Lepid. Soc. 26: 237–244.<br />

MACNEILL, CD. 1963. Callophrys fotis (Strecker) from the San Francisco<br />

Bay area. Pan-Pacific Entomologist 39: 60.<br />

ORSAK, L. and WHITMAN, D.W. 1986. Chromatic color polymorphism in<br />

Callophrys mossii bayensis larvae (<strong>Lycaenidae</strong>): spectral characteristics,<br />

short-term color shifts, and natural morph frequencies. J. Res. Lepid. 25:<br />

188–201.<br />

WEISS, S.B. and MURPHY, D.D. 1991. Thermal microenvironments and the<br />

restoration <strong>of</strong> rare butterfly habitat. In: J. Berger (Ed.) Environmental<br />

Restoration. Island Press, Covelo, California, pp. 50–60.


Country: U.S.A.<br />

The Lotis Blue, Lycaeides idas lotis (Lintner)<br />

R.A. ARNOLD<br />

Entomological Consulting Services Limited, 104 Mountain View Court, Pleasant Hill, CA 94523, U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – probably extinct;<br />

endangered (Red List).<br />

The Lotis Blue was regarded by Arnold (1985) as probably<br />

the rarest resident butterfly in the continental United States.<br />

Historically, it was probably restricted to just a few localised<br />

colonies in coastal northern California. Although it was listed<br />

as an endangered species in 1976, today it is feared to be extinct.<br />

Taxonomy and Description: The Lotis Blue was formerly<br />

considered to be a subspecies <strong>of</strong> Lycaeides argyrognomon<br />

(Bergstrasser), and was listed as endangered under the name L.<br />

argyrognomon lotis. However, European butterfly taxonomists<br />

have recently examined the types <strong>of</strong> L. argyrognomom and L.<br />

idas. They concluded that all North American taxa, which were<br />

formerly called L. argyrognomon, should now be called L. idas.<br />

Hence, the Lotis Blue is now referred to as L. idas lotis. Twelve<br />

subspecies <strong>of</strong> L. idas (Linneaus) have been described from<br />

North America (Downey 1975). The Lotis Blue is one <strong>of</strong> the<br />

larger subspecies <strong>of</strong> L. idas. It is also separable from other<br />

described species by its wing colours and markings.<br />

Distribution: Nearly all collections or sightings <strong>of</strong> the Lotis<br />

Blue since 1933 have been from a single location, near the town<br />

<strong>of</strong> Mendocino in Mendocino County, California (Arnold 1991).<br />

This location is a Sphagnum bog situated in the right-<strong>of</strong>-way for<br />

the Elk-Fort Bragg 60-kV transmission line, operated by the<br />

Pacific Gas and Electric Company (P.G. & E). The bog is about<br />

lha in size and is surrounded by Red Alder (Alnus rubra)<br />

riparian forest, Northern Bishop Pine (Pinus muricata) forest,<br />

and Mendocino Pygmy Cypress (Cupressus pygmaea) forest.<br />

Characteristic understory vegetation includes various ericaceous<br />

shrubs, sedges, and ferns.<br />

Historical records <strong>of</strong> the Lotis Blue reveal that it was known<br />

from only a few other locations, between Point Arena and Fort<br />

Bragg in coastal Mendocino County. Reports <strong>of</strong> the butterfly's<br />

probable occurrence in Sonoma and northern Marin counties by<br />

Tilden (1965) are unsubstantiated by any specimens. Because<br />

<strong>of</strong> substantial differences in the types <strong>of</strong> vegetation that occur<br />

in these areas, it is doubtful that the butterfly ever occurred<br />

143<br />

outside <strong>of</strong> coastal Mendocino County. Fewer than 75 specimens<br />

are housed in North American entomological collections (Arnold<br />

1991).<br />

In 1990, P.G. & E. sponsored an extensive survey <strong>of</strong> the<br />

butterfly and its suspected larval foodplant at 23 locations in<br />

coastal Mendocino County, the historical geographic range <strong>of</strong><br />

the Lotis Blue. Unfortunately, no specimens <strong>of</strong> the Lotis Blue<br />

were found; indeed, none have been seen since 1983 and the<br />

negative results <strong>of</strong> this survey suggest that the butterfly may<br />

now be extinct (Arnold 1991).<br />

Population Size: On June 19th and 20th, 1953, J.W. Tilden<br />

collected at least 26 adults <strong>of</strong> the Lotis Blue from the bog<br />

population <strong>of</strong> the eventual P.G. & E. transmission line, which<br />

are preserved in North American entomological collections.<br />

During 1977–1989, Arnold saw only 26 adult butterflies during<br />

67 days <strong>of</strong> field work at the P.G. & E. location, and none after<br />

1983. Because <strong>of</strong> the small size <strong>of</strong> the butterfly's habitat, its<br />

population numbers probably were never greater than a few<br />

hundred individuals per season at the powerline bog. The<br />

butterfly's very rarity has precluded detailed investigations<br />

about its population size and structure, as have been obtained<br />

for other endangered lycaenids that occur in California (Arnold<br />

1983).<br />

Habitat and ecology: Because <strong>of</strong> the Lotis Blue's rarity, little<br />

is known about its specific habitat requirements and ecology. In<br />

northern California, other subspecific taxa <strong>of</strong> Lycaeides idas<br />

typically occur in wet meadows, bogs, seeps or springs, and<br />

along streamsides. Populations <strong>of</strong> these butterflies are typically<br />

associated with small, and <strong>of</strong>ten isolated, patches <strong>of</strong> their larval<br />

foodplants. Known larval foodplants include legumes that<br />

grow in these wet habitats, in particular Lotus oblongifolius,<br />

Lupinus polyphyllus and Astragalus whitneyi (A. Shapiro, pers.<br />

comm.).<br />

Four legume species have been observed growing in or very<br />

near the bog at the primary Lotis Blue population site along the<br />

transmission line. Of these legumes, Lotus formosissimus (Coast<br />

Trefoil) is the most likely candidate to be the butterfly's larval<br />

foodplant, since it grows in the bog where most specimens <strong>of</strong><br />

the Lotis Blue have been observed. Also, a female was observed


(J.F. Emmel, cited in Arnold 1985) attempting to oviposit on<br />

this plant.<br />

A hypothetical life cycle and natural history <strong>of</strong> the Lotis<br />

Blue can be surmised based on circumstantial evidence from<br />

related taxa whose biologies are better known. Like other taxa<br />

<strong>of</strong> L. idas in northern California, the Lotis Blue is probably<br />

univoltine. Historical collection records indicate that adults<br />

may be present from mid-May through mid-July. Eggs are laid<br />

throughout the adult flight season and newly hatched larvae<br />

probably begin to feed immediately. Partially grown larvae,<br />

probably second instars, diapause until the following spring,<br />

when larval development is completed in about four to six<br />

weeks after feeding resumes. Presumably, the pupal stage lasts<br />

no more than a few weeks.<br />

Threats: Because <strong>of</strong> the butterfly's extremely limited<br />

geographical range, and the small size <strong>of</strong> its only known habitat,<br />

the Lotis Blue is extremely vulnerable to any type <strong>of</strong> habitat loss<br />

or alteration. Collecting <strong>of</strong> specimens could also be detrimental.<br />

Arnold (1985) speculated that drought may have previously<br />

affected the butterfly's habitat by decreasing water levels in the<br />

bog. However, more recent information suggests that<br />

successional changes in the vegetation at the transmission line<br />

site, and at other potential population sites, are probably<br />

responsible for the recently observed decline <strong>of</strong> the butterfly.<br />

The leguminous foodplants <strong>of</strong> other northern California<br />

populations <strong>of</strong> L. idas generally grow in small patches in wet<br />

habitats that are at the early stages <strong>of</strong> vegetation succession.<br />

Circumstantial evidence from the 1990 P.G. & E. sponsored<br />

study suggests that Lotus formosissimus, the suspected larval<br />

foodplant <strong>of</strong> the Lotis Blue, also grows in greatest abundance in<br />

the early successional stages <strong>of</strong> wet areas, such as bogs, plus the<br />

headwaters and shorelines <strong>of</strong> streams (Arnold 1991).<br />

Furthermore, examination <strong>of</strong> field notes from cadastral surveys<br />

and maps prepared by the U.S. Coast and Geodetic Survey,<br />

during the late 1800s and early 1900s indicate that the<br />

transmission line site and much <strong>of</strong> the coastal area within the<br />

butterfly's historic range were logged then. Indeed, many areas,<br />

including the transmission line bog, which today support dense<br />

forest vegetation, were open fields a century ago. Thus, the<br />

forests surrounding the transmission line bog represent regrowth<br />

rather than natural habitat. Comparison <strong>of</strong> a series <strong>of</strong> aerial<br />

144<br />

photos covering the past several decades depict how rapidly the<br />

vegetation has changed from a more open area to a dense forest<br />

with a closed canopy in many places. Presumably, as the<br />

successional changes have proceeded, the increasing density <strong>of</strong><br />

the vegetation gradually choked out the suspected larval<br />

foodplant <strong>of</strong> the butterfly, which prefers open areas. By 1990,<br />

only 15 specimens <strong>of</strong> L. formosissimus were observed at two<br />

locations in or near the transmission line bog (Arnold 1991),<br />

and only two specimens were observed growing in the bog<br />

during 1992 (Arnold, unpublished data).<br />

<strong>Conservation</strong>: Clearly, basic ecological and natural history<br />

information about the butterfly is needed to identify potential<br />

habitat sites within its historic range and to manage these areas<br />

to benefit the Lotis Blue. Confirmation <strong>of</strong> the butterfly's larval<br />

foodplant is essential to identify its breeding habitat. Because<br />

<strong>of</strong> the butterfly's dubious status, surveys should be undertaken<br />

to determine if it even still exists.<br />

Objectives <strong>of</strong> the butterfly's recovery plan (Arnold 1985)<br />

are to: (1) protect the butterfly and its habitat at the only known<br />

site; (2) establish three new viable populations at different sites;<br />

and (3) determine the extent <strong>of</strong> the population and size <strong>of</strong> secure<br />

habitats needed so the butterfly can be reclassified as 'threatened'<br />

rather than 'endangered'. However, until the basic biological<br />

information about the butterfly is obtained, the objectives <strong>of</strong> the<br />

recovery plan probably cannot be achieved.<br />

References<br />

ARNOLD, R.A. 1983. Ecological studies <strong>of</strong> six endangered butterflies<br />

(Lepidoptera: <strong>Lycaenidae</strong>): island biogeography, patch dynamics, and<br />

the design <strong>of</strong> habitat preserves. Univ. Calif. Publns Entomol. 99: 1–161.<br />

ARNOLD, R.A. 1985. (Main author <strong>of</strong>) U.S. Fish & Wildlife Service. Lotis<br />

Blue butterfly recovery plan. U.S. Fish & Wildlife Service. Portland,<br />

Oregon. 46pp.<br />

ARNOLD, R.A. 1991. Biological studies <strong>of</strong> the endangered Lotis Blue<br />

butterfly for P.G. & E. 's Elk-Fort 60kV transmission line. Pacific Gas &<br />

Electric Company. San Ramon, California. 46pp. and appendices.<br />

DOWNEY, J.C. 1975. Genus Plebejus Kluk. In: Howe, W.H. (Ed.). The<br />

<strong>Butterflies</strong> <strong>of</strong> North America. Doubleday. Garden City, New York. pp.<br />

337–350.<br />

TILDEN, J.W. 1965. <strong>Butterflies</strong> <strong>of</strong> the San Francisco Bay Region. Univ. <strong>of</strong><br />

Calif. Press. Berkeley, California. 88pp.


Additional taxa <strong>of</strong> concern in southern California<br />

Three additional lycaenid butterflies <strong>of</strong> conservation interest<br />

occur in southern California:<br />

The Human Folly Blue, Philotes<br />

sonorensis extinctis Mattoni<br />

The subspecies occurred across a l000ha site in the upper San<br />

Gabriel river wash. It has been extinct since 1968, having been<br />

eliminated by the activities <strong>of</strong> the U.S. Army Corps <strong>of</strong> Engineers<br />

to provide a spreading basin for subsurface water recharge.<br />

Ironically the Corps is today charged with the responsibility <strong>of</strong><br />

protecting wetlands and species. Details <strong>of</strong> this subspecies can<br />

be found in Mattoni (1991).<br />

Reference<br />

MATTONI, R.H.T. 1991. An unrecognised, now extinct, Los Angeles area<br />

butterfly (<strong>Lycaenidae</strong>). J. Res. Lepid. 28: 297–309 (1989).<br />

The Santa Monica Mountains Hairstreak,<br />

Satyrium auretorum fumosum Emmel &<br />

Mattoni<br />

This hairstreak is only known from a few localities within a<br />

circumscribed area <strong>of</strong> about 25,000ha in the western Santa<br />

Monica mountains. The range is completely surrounded by<br />

urban development and itself is being further fragmented by<br />

developments. Battlelines between environmentalists and<br />

developers have been drawn to define the future <strong>of</strong> the mountains.<br />

The compromises finally reached regarding habitat preservation<br />

<strong>of</strong> the area may serve as a model for the rest <strong>of</strong> the United States.<br />

With land values in the order <strong>of</strong> one million dollars per hectare,<br />

economic pressures even affect biologist consultants. An effort<br />

to federally list the species is underway.<br />

R.H.T. MATTONI<br />

9620 Heather Road, Beverly Hills, California 90210, U.S.A.<br />

145<br />

The San Gabriel Blue, Plebejus saepiolus<br />

undescribed subspecies<br />

This is another extinct taxon. Known only from a few meadows<br />

in the Big Pine recreation area on the north slope <strong>of</strong> the San<br />

Gabriel mountains it was last seen in the mid 1980s. Its<br />

distribution was very limited, since its wet meadow habitat,<br />

necessary to support the clover foodplant, is very restricted<br />

across the dry north slope. Extinction was brought on by<br />

draining <strong>of</strong> the limited meadow habitats by the U.S. Forestry<br />

Service.


Selected Neotropical species<br />

K.S. BROWN, JR.<br />

Departamento de Zoologia, Instituto de Biologia, Universidade Estadual de Campinas, C.P. 6109, Campinas, São Paulo,<br />

13.081, Brazil<br />

Styx infernalis Staudinger<br />

Country: Peru.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

This species is probably the most primitive <strong>of</strong> the Riodininae,<br />

confined to a very small region <strong>of</strong> high species diversity and<br />

endemicity. It is very rarely observed.<br />

Taxonomy and Description: A medium-sized, dirty transparent<br />

grey butterfly with narrow wings and a heavy black body,<br />

seeming rather like a Geometrid or Lymantriid moth. It has<br />

short antennae.<br />

Distribution: S. infernalis is known only from central and<br />

southern Peru, at elevations between 1000 and 1600m (Figure 1).<br />

Population Size: Not known.<br />

Habitat and Ecology: This species inhabits primeval cloud<br />

forest with steep slopes and torrents. It is active in sun patches<br />

near midday, with a weak, almost gliding flight, in small<br />

aggregations near rushing streams. The early stages are unknown.<br />

Threats: Habitat conversion for c<strong>of</strong>fee or other plantations<br />

(very scattered colonies).<br />

<strong>Conservation</strong>: The main need is to locate colonies and secure<br />

large tracts <strong>of</strong> undisturbed cloud forest.<br />

Nirodia belphegor (Westwood)<br />

Country: Brazil.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

Nirodia is a monotypic genus (very close toRhetus, composed<br />

<strong>of</strong> common tropical species), whose only species is confined to<br />

high-altitude rockfields in a very ancient environment. It has<br />

146<br />

been observed fewer than ten times since its original description<br />

140 years ago in spite <strong>of</strong> extensive human activity in the region.<br />

Taxonomy and Description: N. belphegor is a medium-small,<br />

yellow-spotted, dark blue butterfly with pointed forewings and<br />

a short broad tail on the hindwings (Figure 1d <strong>of</strong> regional<br />

account), strongly reminiscent <strong>of</strong> Rhetus periander (Cramer). It<br />

seems quite close to a fossil riodinine from the Eocene (Durden<br />

and Rose 1978).<br />

Distribution: It is known from only four localities in the Serras<br />

do Espinhago and Cipó in central Minas Gerais, Brazil, at<br />

elevations above 1000m (Figure 1) where it is seen very<br />

sporadically.<br />

Population Size: Not known.<br />

Habitat and Ecology: It inhabits 'campo rupestre', a xeric<br />

system on rocky soils at about 1000m elevation, characterised<br />

by strong endemism in plants (Velloziaceae, Eriocaulaceae,<br />

Xyridaceae) and animals (including amphibians), with ancient<br />

affinities to similar systems in Africa. Males are seen resting on<br />

the ground with wings outspread (like Rhetus) drinking water<br />

beside small creeks rushing down through rockfields. They<br />

make short sallies out from perching places on sunlit rocks and<br />

dart irregularly about as if defending a territory in that space<br />

(fide Ivan Sazima). Early stages are unknown, as are any further<br />

biological details on the species.<br />

Threats: The 'campo rupestre' system, restricted in area, is<br />

under very heavy pressure from removal and exportation <strong>of</strong><br />

dried plants, and is subject to extensive burning and trampling.<br />

<strong>Conservation</strong>: The species should be preserved in the recently<br />

declared Serra do Cipó National Park, and the proposed Caraça<br />

Natural Park.<br />

Reference<br />

DURDEN, C.J. and ROSE, H. 1978. <strong>Butterflies</strong> from the middle Eocene: the<br />

earliest occurrence <strong>of</strong> Fossil Papilionoidea (Lepidoptera). Pearce-Sellards<br />

Series, Texas Memorial Museum 29: 1–25.


Figure 1. Distribution (A–E) <strong>of</strong> selected Neotropical species <strong>of</strong> <strong>Lycaenidae</strong>.<br />

Joiceya praeclarus Talbot<br />

Country: Brazil.<br />

Status and <strong>Conservation</strong> Interest: Status – endangered.<br />

Joiceya is a monotypic genus known only from the type<br />

series, from a small, heavily collected and increasingly converted<br />

region.<br />

147<br />

Taxonomy and Description: A 'Setabis alcmaeon (Lewis<br />

1973, plate 77: Figure 5) -like' pattern: small, with a pointed<br />

forewing and elongated hindwing; dull grey-brown underside<br />

with faint distal lines; strong blue upperside with a black baseto-margin<br />

line on the hindwing; and a black area between<br />

forewing base and submarginal spots. It is like no other species<br />

or genus.


Distribution: J. praeclarus is known only from an original<br />

collection in central Mato Grosso (Cuiabá and Tombador),<br />

Brazil (Figure 1). It has not been collected in intensive recent<br />

work in the region.<br />

Population Size: Not known.<br />

Habitat and Ecology: The ecology <strong>of</strong> this species is not<br />

known, but it probably inhabits the understorey <strong>of</strong> isolated<br />

humid headwater or spring-fed forests in strongly scarped<br />

regions within the cerrado landscape (chapadas).<br />

Threats: The area is being used intensively for colonisation,<br />

ranching, hydroelectric projects, mines, and industrial farms,<br />

with removal <strong>of</strong> all original vegetation in parts.<br />

<strong>Conservation</strong>: The species may be present in the new Chapada<br />

dos Guimarães National Park and other small reserves in the<br />

region, but it needs to be relocated and colonies specifically<br />

saved.<br />

Reference<br />

LEWIS, H.L. 1973. <strong>Butterflies</strong> <strong>of</strong> the World. Follett, Chicago.<br />

Arcas, five rarer species<br />

Countries: Panama, Ecuador, Peru, Brazil.<br />

Status and <strong>Conservation</strong> Interest: Status – rare or vulnerable<br />

(with the exception <strong>of</strong> A. imperialis which is not considered<br />

threatened).<br />

These are the most exquisite <strong>of</strong> all neotropical Theclinae,<br />

typical <strong>of</strong> large areas <strong>of</strong> virgin wet forest and usually disappearing<br />

in disturbed areas. They are easy to find when present and are<br />

thus good indicators <strong>of</strong> undisturbed forest systems.<br />

Taxonomy and Description: These are largish, two-tailed<br />

species with an additional well-separated long anal lobe <strong>of</strong> the<br />

hindwing (thus flashing six mobile 'antennae' to the false<br />

head). They are brilliant striated green on the ventral surface<br />

(<strong>of</strong>ten with additional nuances <strong>of</strong> rose and chrome yellow) and<br />

iridescent blue or green above. The genus was revised by<br />

Nicolay (1971). Illustrations <strong>of</strong> at least A. imperialis appear in<br />

most popular butterfly books, and A. ducalis has been called the<br />

most beautiful small butterfly in the neotropics.<br />

Distribution: The non-threatened A. imperialis (Cramer) occurs<br />

from Mexico to southern Brazil, with northerly females<br />

sometimes showing much rose colour ventrally. A. cypria<br />

(Geyer) is infrequent from Mexico to western Colombia, and A.<br />

ducalis (Westwood) is very local in southeastern Brazilian<br />

mountains. The A. tuneta superspecies (delphia Nicolay in<br />

148<br />

Costa Rica, and tuneta (Hewitson) from Peru to southeastern<br />

Brazil) are very rarely encountered. A. jivaro Nicolay is known<br />

only from the two types in a single locality, in eastern Ecuador<br />

and A. splendor (Druce), not seen for 110 years after its original<br />

collection, is restricted to scattered cloud forest localities from<br />

Costa Rica to W. Ecuador (Figure 1). Further taxa may still be<br />

found.<br />

Population Size: Not known.<br />

Habitat and Ecology: Arcas males are most <strong>of</strong>ten seen in early<br />

afternoon on forested hilltops or ridges, where they sit high<br />

(5–10m) in the trees in the sun, looking out from leaves over a<br />

green sunlit space. They vigorously defend this space against<br />

other males while waiting for a female to arrive. Mated females<br />

are found lower down, inspecting growing tips <strong>of</strong> the foodplants<br />

(Annonaceae: Rollinia, and Lauraceae for A. ducalis). Two or<br />

three species can be found on a single hilltop in Panama or<br />

southeastern Brazil, but such favoured sites are very rarely<br />

found. Flower visiting is infrequent, mostly before midday, but<br />

may be prolonged. Flight is very rapid, a brilliant green swirl<br />

accompanied by blue flashes, with the insect returning to the<br />

same or a nearby perch. Most populations occur throughout the<br />

year but the insects are commoner on sunny days in the rainier<br />

seasons.<br />

Threats: Moderate modification <strong>of</strong> the habitat may eliminate<br />

male perching sites and prevent the sexes from meeting and<br />

mating, in the sparse populations <strong>of</strong> these species. Two <strong>of</strong> the<br />

species are known from very few localities and may easily be<br />

eliminated.<br />

<strong>Conservation</strong>: For the rarest <strong>of</strong> these species, A. jivaro and A.<br />

splendor, further localities should be looked for and protected<br />

Arcas ducalis (photo by K.S. Brown, Jr.).


wherever possible. Three <strong>of</strong> the more widespread species<br />

(cypria, ducalis, and tuneta), though known from more areas,<br />

are still very rarely encountered. Along with the more frequently<br />

encountered A, imperialis, they should be reported and monitored<br />

as good indicators <strong>of</strong> the health <strong>of</strong> intact, rich tropical wet forest<br />

ecosystems.<br />

Reference<br />

NICOLA Y, S.S. 1971. A review <strong>of</strong> the genus Arcas with descriptions <strong>of</strong> new<br />

species (<strong>Lycaenidae</strong>, Strymonini). J. Lepid. Soc. 25: 87–108.<br />

Arawacus aethesa (Hewitson)<br />

Country: Brazil.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

This species is endemic to wet forests in a restricted area <strong>of</strong><br />

the Atlantic lowlands <strong>of</strong> southeastern Brazil. Since the building<br />

<strong>of</strong> a major new highway twenty years ago, over 90% <strong>of</strong> the<br />

vegetation has been cut over, greatly disturbed or removed.<br />

Taxonomy and Description: A. aethesa is closely related to A.<br />

aetolus (= linus Auctt.), the classical 'false head' phenotype<br />

which is very widespread in the neotropics, with convergent<br />

black lines on the cubital spot <strong>of</strong> the hindwing margin, from the<br />

forewing costa. It is easily distinguished by its smoky brown<br />

undersurface, with a yellow submarginal band.<br />

Distribution: It is known only from southern Bahia, northern<br />

Espírito Santo, and eastern Minas Gerais, in the Atlantic forests<br />

<strong>of</strong> eastern Brazil.<br />

Population Size: Not known.<br />

Habitat and Ecology: It has been observed in sunlit columns<br />

in the interior <strong>of</strong> wet lowland forest. The foodplant is almost<br />

surely Solarium leaves.<br />

Threats: Habitat reduction is the primary threat to this species<br />

and many others endemic to Atlantic lowland forest.<br />

<strong>Conservation</strong>: The species exists in several reserves within its<br />

range, but needs to be specially protected as it uses secondarysuccession<br />

plants in a forest biome.<br />

149<br />

Cyanophrys bertha (Jones)<br />

Country: Brazil.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable.<br />

A very rarely seen (only the single type existed among<br />

collections in all American and European museums until very<br />

recently) and spectacular species typical <strong>of</strong> large, species-rich,<br />

high-elevation sites in the coastal mountains, where many other<br />

unusual and little-known species occur.<br />

Taxonomy and Description: Like other Cyanophrys, it is blue<br />

above and pea-green below, but in contrast to them it has a white<br />

stripe across both wings on the underside, bordered basally with<br />

diffuse blue-white blotches.<br />

Distribution: The species is known in Brazilian collections<br />

from over 1000m elevation in the mountains <strong>of</strong> Minas Gerais<br />

(Pocos de Caldas, Barbacena), Rio de Janeiro (Petr6polis), Sao<br />

Paulo (Serra do Japi, where quite frequent at times), and Parana<br />

from a total <strong>of</strong> fewer than 20 specimens.<br />

Population Size: Not known.<br />

Habitat and Ecology: Males frequent hilltops in the early<br />

afternoon, where they perch in the crowns (not usually the<br />

highest points) <strong>of</strong> tall scraggly trees, changing perches every<br />

few minutes or when approached. Both sexes can be found on<br />

flowers in the morning, and females frequent sunny glades on<br />

hillsides. Early stages are still unknown. It is associated with C.<br />

herodotus (F.), C. amyntor (Cramer), C. acaste (Prittwitz), C.<br />

remus (Hewitson) and C. nr. pseudolongula (Clench), the last<br />

three <strong>of</strong>ten perching on the same high-elevation hilltop trees.<br />

Threats: Like many highly localised and rarely seen Theclinae,<br />

it can be eliminated from the few local colonies by minor habitat<br />

alteration.<br />

<strong>Conservation</strong>: The species should be preserved in a number <strong>of</strong><br />

parks and inaccessible areas.


Neotropical <strong>Lycaenidae</strong> endemic to high elevations in SE Brazil<br />

K.S. BROWN, JR.<br />

Departamento de Zoologia, Instituto de Biologia, Universidade Estadual de Campinas, C.P. 6109, São Paulo, 13. 081, Brazil<br />

Country: Brazil.<br />

Status and <strong>Conservation</strong> Interest: Status – threatened<br />

community.<br />

These are rarely observed, <strong>of</strong>ten very specialised species<br />

and genera, in very local habitats in a strongly heterogeneous<br />

landscape. The area is subject to intensive human activity and<br />

there is considerable modification <strong>of</strong> the vegetation.<br />

Taxonomy and Description: In the Riodininae the community<br />

includes: three monotypicgenera:Nirodia belphegor (discussed<br />

under 'Selected Neotropical Species'); Eucorna sanarita<br />

(Schaus), dark with a 'fuzzy' pattern; Petrocerus catiena<br />

(Hewitson), similar to a dark Calydna; isolated species <strong>of</strong><br />

Mesosemia (?) (M.? acuta Hewitson, beige with a falcate<br />

forewing); Calydna (a still undescribed species, very small<br />

with orange spots and a falcate forewing); Crocozona (?) (C.?<br />

croceifasciata Zikán, small with a transverse orange band<br />

across both wings); a species <strong>of</strong> Mesenopsis (M. albivitta<br />

(Lathy), imitating common Dioptid moths with orange bars on<br />

each wing (see next account); Panara ovifera Seitz; Mycastor<br />

leucarpis (Stichel); Argyrogrammana caesarion Rebillard and<br />

a number <strong>of</strong> species <strong>of</strong> Napaea (all dark brown).<br />

Endemic theclines are many, though not equal to the highaltitude<br />

groups in the Andes; some are more 'furry' than their<br />

lower-altitude congeners (due to the cold?) and many have<br />

rather sombre patterns (but see Cyanophrys bertha, Brown, this<br />

volume).<br />

150<br />

Distribution: The above groups are found in the Serras do Mar<br />

and da Mantiqueira, from the interior <strong>of</strong> Bahia and Espírito<br />

Santo south to Santa Catarina and northern Rio Grande do Sul.<br />

They are most common in very high areas in Rio de Janeiro/<br />

Minas Gerais/São Paulo.<br />

Population Sizes: Not known.<br />

Habitat and Ecology: The species fly rarely, but have been<br />

noted flying in sunny weather near midday, in variable seasons<br />

and habitats. Most have been found too sporadically to make<br />

any generalisations, but as a rule they are likely to be in wet<br />

habitats or bamboo forests, not rare when found, occasionally<br />

partial to hilltops, and probably tightly associated with special<br />

host plants or ants.<br />

Threats: The exceedingly heterogeneous habitats are full <strong>of</strong><br />

small micro-islands <strong>of</strong> different vegetation (<strong>of</strong>ten no more than<br />

a few hundred square metres on a hilltop, hillside or gully)<br />

whose destruction leads to extinction <strong>of</strong> the populations.<br />

<strong>Conservation</strong>: Although many large, inaccessible areas still<br />

exist in these regions, some <strong>of</strong> them <strong>of</strong>ficially preserved, little<br />

information is available about the presence <strong>of</strong> these rare species,<br />

and possibly other <strong>Lycaenidae</strong> still undescribed. Much<br />

exploration is needed to establish the localities <strong>of</strong> colonies and<br />

preserve them.


Country: Brazil.<br />

Riodininae: Amazonian genera with most species<br />

very rare or local<br />

K.S. BROWN, JR.<br />

Departamento de Zoologia, Institute de Biologia, Universidade Estadual de Campinas, C.P. 6109,<br />

Säo Paulo, 13. 081, Brazil<br />

Status and <strong>Conservation</strong> Interest: Status – threatened<br />

community.<br />

There are a number <strong>of</strong> forest species that are very rarely<br />

seen, concentrated in compact taxonomic groups, whose local<br />

colonies, <strong>of</strong>ten very far apart, may be eliminated easily by<br />

modest alteration <strong>of</strong> the habitat.<br />

Description: The group includes:<br />

• four species <strong>of</strong> Alesa (telephae Boisduval, fournierae<br />

Rebillard, neagra Röber, thelydryas Bates, known from<br />

very few specimens but not A. prema Godart or A. amesis<br />

Cramer),<br />

• Mimocastnia rothschildi Seitz (a hypertrophic representative<br />

<strong>of</strong> the same phenotype and lineage);<br />

• both species <strong>of</strong> Colaciticus;<br />

• three species <strong>of</strong> Mesenopsis (a fourth is montane in SE<br />

Brazil, see preceding account);<br />

• all species <strong>of</strong> Xenandra including X. pulcherrima (Herrich-<br />

Schäffer) (usually placed in Melanis);<br />

• most species <strong>of</strong> Esthemopsis and Symmachia;<br />

• Zelotaea phasma Bates;<br />

151<br />

• all species <strong>of</strong> Dysmathia.<br />

The species vary from uniformly dingy (the last genus) to<br />

uniformly white (penultimate), with many having strong colours<br />

<strong>of</strong> yellow, orange and red or green; all are small except for<br />

Mimocastnia.<br />

Population Sizes: Not known.<br />

Habitat and Ecology: They are typically found as isolated<br />

males in very high-diversity hilltops or clearings, very<br />

occasionally on flowers or in small assemblages. These species<br />

all inhabit the deep forest; some may be canopy dwellers, but<br />

most just seem to be very rare and sporadic in occurrence,<br />

presumably due to excessively narrow ecological tolerances.<br />

Threats: The very rarefied distributions indicate that the<br />

conversion <strong>of</strong> small areas may eliminate local colonies which<br />

will not be re-established in nearby intact habitat.<br />

<strong>Conservation</strong>: Wherever colonies are known, they should be<br />

protected by moderate-sized reserves which can maintain natural<br />

processes and heterogeneity.


Theclinae endemic to the Cerrado vegetation (central Brazil)<br />

Country: Brazil.<br />

K.S. BROWN, JR.<br />

Departamento de Zoologia, Institute de Biologia, Universidade Estadual de Campinas,<br />

C.P. 6109, Sao Paulo, 13. 081, Brazil<br />

Status and <strong>Conservation</strong> Interest: Status – threatened<br />

community.<br />

These are very local species that are indicators <strong>of</strong> healthy<br />

and diverse, well-watered cerrado habitats which contain the<br />

full range <strong>of</strong> microenvironments and successional vegetation<br />

typical <strong>of</strong> the region.<br />

Taxonomy and Description: Coming from many groups and<br />

genera in the Theclinae, most <strong>of</strong> the characteristic 'cerrado<br />

species' have large red markings or blotches on the underside,<br />

almost as if they formed an environment-mediated mimicry<br />

ring. Typical representatives are Arawacus tarania (Hewitson),<br />

Strymon tegaea (Hewitson), S. ohausi (Spitz), S. sp. (zibagroup),<br />

'Thecla' mantica Druce, Magnastigma julia Nicolay,<br />

'Thecla' socia Hewitson, and 'Thecla' bagrada Hewitson. All<br />

are distinctive and restricted to the cerrado domain and its<br />

peripheries.<br />

Distribution: Central Brazil Plateau in Goiás, Distrito Federal,<br />

Minas Gerais, Mato Grosso, Mato Grosso do Sul, and parts <strong>of</strong><br />

Bahia, Sao Paulo and Parana, in the cerrado.<br />

Population Sizes: Not known.<br />

152<br />

Habitat and Ecology: The habitats <strong>of</strong> the group are variable:<br />

S. tegaea prefers wet grasslands by headwater woods, while the<br />

minute S. ohausi is restricted to tiny grassy marshes in small<br />

sinkholes within the cerrado. A. tarania is more widespread in<br />

scrub with a grassy understorey, the juveniles feeding on small<br />

Leguminosae (K. Ebert, pers. comm.); 'Th.' mantica and M.<br />

julia prefer bushy cerrado, the larvae <strong>of</strong> the former living on<br />

Chrysobalanaceae (K. Ebert, pers. comm.). The other three<br />

'Thecla', are <strong>of</strong>ten found perched in medium-sized trees, near<br />

the end <strong>of</strong> the afternoon. Most are scarce.<br />

Threats: Large areas <strong>of</strong> cerrado are being occupied by intensive,<br />

mechanised and chemical agriculture, completely destroying<br />

and poisoning the diverse and complex natural systems. As<br />

several <strong>of</strong> these species are intensely localised and occupy rare,<br />

scattered and very specific habitats, their few local populations<br />

are especially vulnerable.<br />

<strong>Conservation</strong>: A number <strong>of</strong> preserved areas in the cerrado<br />

region probably include the commoner species <strong>of</strong> Theclinae<br />

endemic to this biome, but colonies <strong>of</strong> the rarer ones such as S.<br />

ohausi and M. julia need to be localised and specifically<br />

protected. The latter has not been seen since the 1969 collecting<br />

trip that led to its discovery and description, when it was found<br />

in only three localities, all <strong>of</strong> precarious preservation today.


Neotropical Riodininae endemic to the Chocó region <strong>of</strong> western<br />

Colombia<br />

Country: Colombia.<br />

C.J. CALLAGHAN<br />

Louis Berger International Inc., 100 Halsted Street, P.O. Box 270, East Orange, N.J. 07019, U.S.A.<br />

Status and <strong>Conservation</strong> Interest: Status – threatened<br />

community.<br />

The area supports a high number <strong>of</strong> rare, specialised species<br />

<strong>of</strong> riodinine butterflies, all very sensitive to alterations in<br />

vegetation patterns. The area is currently under pressure from<br />

human activity, especially logging <strong>of</strong> pulp-wood for paper<br />

mills.<br />

Taxonomy and Description: The fauna is taxonomically poorly<br />

known, so that future study may well reveal additional species<br />

or genera which are endemic to the region. Preliminary work by<br />

Callaghan (1985) suggests that 33 (36%) <strong>of</strong> a total <strong>of</strong> 91 species<br />

<strong>of</strong> the known Riodininae are endemic. Among these areEuselasia<br />

rhodogyne (Godman), E. violacea Lathy, two undescribed<br />

Euselasia, Mesosemia asa iphigenia Stichel, M. sibyllina<br />

Staudinger, Eurybia juturna cyclopia Stichel, Lucillella sp.<br />

nov., Calospila rubrica (Stichel), C. asteria (Stichel), C. caligata<br />

(Stichel), Nymphidium balbinus Staudinger, and Stalachtis<br />

magdalenae cleove Staudinger. In addition to these endemics,<br />

the Chocó above 1000m also forms the last refuge for Mesosemia<br />

mehida Hewitson and M. bifasciata Hewitson, described from<br />

western Ecuador.<br />

Distribution: The Chocó region extends from north <strong>of</strong> Quibdó<br />

south along the western slopes <strong>of</strong> the Cordillera Occidental to<br />

northwestern Ecuador. The fauna shows strongest affinities<br />

with Panama/Costa Rica, with which it shares 58% <strong>of</strong> its<br />

Riodininae. Because <strong>of</strong> this shared influence from Panama,<br />

43% <strong>of</strong> the Chocó fauna is also found in the Cauca Valley to the<br />

east <strong>of</strong> the Cordillera; but from there eastwards, the faunal<br />

similarities drop rapidly. Only 17% <strong>of</strong> the taxa are also<br />

encountered on the eastern (Amazonian) slope <strong>of</strong> the Cordillera<br />

Oriental.<br />

153<br />

Population Sizes: Not known.<br />

Habitat and Ecology: The characteristic <strong>of</strong> the Chocó is its<br />

high rainfall, as much as 13 metres a year east <strong>of</strong> Quibdó. The<br />

rain is nearly constant, though slightly less in January-February<br />

and July-August, and <strong>of</strong>ten mostly in the afternoon and evening.<br />

The butterflies take to the wing during short intervals <strong>of</strong> sun,<br />

usually in the morning and early afternoon. Most species<br />

concentrate on hilltops, although many are found in the forests<br />

along trails and streams, <strong>of</strong>ten deeply cut into the weathered<br />

soil.<br />

Threats: Nearly all species are very sensitive to habitat<br />

alteration, particularly cutting <strong>of</strong> the rain forest. A significant<br />

fall in the number <strong>of</strong> hilltopping species has been seen as a<br />

function <strong>of</strong> the cutting <strong>of</strong> larger trees on the slopes.<br />

<strong>Conservation</strong>: Reserves in the Chocó are few and only<br />

established with great difficulty due to the conflicting economic<br />

interests, particularly from the paper industry. However, due to<br />

the uniqueness <strong>of</strong> the habitat and its high endemism, combined<br />

with the rudimentary level <strong>of</strong> existing knowledge, every effort<br />

should be made to support the investigation and establishment<br />

<strong>of</strong> sustainable reserves in the Chocó.<br />

Reference<br />

CALLAGHAN, C.J. 1985. Notes on the zoogeographic distribution <strong>of</strong><br />

butterflies in the subfamily Riodininae in Colombia. In: Proc. 2nd Symp.<br />

Neotropical Lepidoptera. J. Res. Lep. Supplement 1: 50–69.


Aloeides dentatis dentatis (Swierstra), Aloeides dentatis maseruna<br />

(Riley); Subfamily Theclinae, Tribe Aphnaeini<br />

Country: South Africa.<br />

S.F. HENNING 1 , G.A. HENNING 2 and M.J. SAMWAYS 3<br />

1 5 Alexander St., Florida 1709, South Africa<br />

2 17 Sonerend St., Helderdruin 1724, South Africa<br />

3 Department <strong>of</strong> Zoology and Entomology, University <strong>of</strong> Natal, Pietermaritzburg 3200, South Africa<br />

Status and <strong>Conservation</strong> Interest: Status – A. d. dentatis:<br />

rare; A. d. maseruna: indeterminate; rare (Red List).<br />

These two subspecies are highly localised endemics from<br />

the central grassland savannah <strong>of</strong> southern Africa. Earlier<br />

recorded sites for both butterflies have been damaged, with<br />

localised loss <strong>of</strong> the species.<br />

Taxonomy and Description. The upperside <strong>of</strong> A. d. dentatis is<br />

orange, with narrow black margins and apical patches (Murray<br />

1935, Pennington 1978). The basal half <strong>of</strong> the forewing costa is<br />

orange. The underside <strong>of</strong> the hindwing is crimson-red with<br />

silvery white and black markings. The diagnostic feature is a<br />

medial series <strong>of</strong> small dentate markings with black along their<br />

outer edge. There is also a form with a pale brown rather than<br />

crimson ground colour, and with indistinct markings. The<br />

female is similar to the male but has more rounded wings.<br />

Forewing lengths: male 14–18mm; female 15–19mm.<br />

A. d. maseruna has a pale tawny-orange upperside with a<br />

dark grey border. The underside <strong>of</strong> the hindwing is pale brown<br />

or pinkish-red with the silvery dentate band more extensive<br />

than in the nominate subspecies. Forewing lengths: male<br />

13.5–18.5mm; female 15–19mm.<br />

Distribution. A. d. dentatis occurs on the highveld <strong>of</strong> the<br />

Transvaal, at Waterval Onder, Ruimsig, Pretoria, Springs,<br />

Alberton and Suikerbosrand.<br />

A. dentatis maseruna was originally recorded from Maseru<br />

in Lesotho. Other localities in the Orange Free State and the<br />

Western Transvaal are now also known.<br />

Population Size: From mark-recapture data collected at the<br />

Ruimsig Reserve in 1989/1990, it was estimated that the<br />

population size <strong>of</strong> A. d. dentatis was 400 specimens on the wing.<br />

The population size <strong>of</strong> other populations <strong>of</strong> A. d. dentatis is not<br />

known. There is no information available on the sizes <strong>of</strong> the<br />

known colonies <strong>of</strong> A. dentatis maseruna.<br />

Habitat and Ecology: A. d. dentatis occurs in highveld<br />

grassland. The adults do not range far from the host plants and<br />

154<br />

ants. The males maintain small territories, on sandy patches<br />

among the foodplants, in which they can be found throughout<br />

most <strong>of</strong> the day. The females fly randomly throughout the area,<br />

and are as common as the males. During partly cloudy days, A.<br />

d. dentatis has been seen to bask on a rock by lying sideways.<br />

It feeds on a variety <strong>of</strong> flowers, and its foodplant at Ruimsig<br />

is Hermannia depressa (Sterculiaceae) at an elevation <strong>of</strong> 1500m.<br />

The eggs are laid in pairs on the underside <strong>of</strong> the leaves <strong>of</strong> the<br />

foodplant. In the Suikerbosrand Nature Reserve, at 1900m, the<br />

foodplant is Lotononis erianthe (Fabaceae), and the eggs are<br />

laid on the stems. At Ruimsig, the larvae shelter during the day<br />

in the nest <strong>of</strong> the widespread ant Acantholepis capensis Mayr.<br />

The larvae apparently release a pheromone which appears to<br />

mimic the brood pheromone <strong>of</strong> the ant, causing the ant to treat<br />

the larvae as its brood. At night, the larvae emerge from the nest<br />

to feed on the foodplant. During these journeys, they apparently<br />

release a pheromone which imitates the alarm pheromone <strong>of</strong> the<br />

ant. This excites the attendant ants and partially protects the<br />

larvae while they feed. The larvae pupate within the ants' nests.<br />

When they emerge from the pupae, the adults run along the<br />

tunnels in the nest with wings still unexpanded, only expanding<br />

them once outside the nest.<br />

The eggs <strong>of</strong> A. d. dentatis are 0.8mm in diameter and 0.5 mm<br />

high. They are bun-shaped, with a bold network <strong>of</strong> ridges. The<br />

colour is creamy-white at first, becoming purple-brown with<br />

development. All six larval instars are similar in appearance.<br />

The head is dark brown with light brown setae. The broad neckshield<br />

and small rounded anal shield are dark brown. The body<br />

is grey with longitudinal streaks and markings <strong>of</strong> reddishbrown.<br />

The tubercle cases on the eighth segment are black and<br />

bear the characteristic protective spines. The retractile tubercles<br />

are white. The honey gland is absent. Maximum length <strong>of</strong> the<br />

final instar larva is 18–19mm. The pupa is 12–13mm long,<br />

golden-brown and rounded (Henning 1983a,b, 1984a,b, 1987;<br />

Tite and Dickson 1968a,b).<br />

Adults are on the wing from August to April with a peak<br />

from October to December (Henningl983a,b, 1984a, 1987;<br />

Tite and Dickson 1968b).<br />

A. d. maseruna inhabits flat grassveld, usually near water or<br />

marshy areas. The habitats are sparsely grassed, the ground<br />

being sandy with gravel. The adults sit on sand or gravel


patches, the males establishing territories from which to chase<br />

intraspecific intruders and court females. The males fly low and<br />

fast, and in wide circles, generally returning to their original<br />

spots. The females fly slowly about in the same areas. When a<br />

female enters the territory, the male approaches from the rear<br />

and 'shimmies' just behind her. When unresponsive, she flies<br />

<strong>of</strong>f. Alternatively, when responsive, she immediately settles.<br />

The male settles beside her and sidles up, but if she is not ready<br />

to mate, she turns away from him in a 'rejection posture',<br />

causing him to fly <strong>of</strong>f. The actual mating has not yet been<br />

recorded. The fertilised female then spends her time searching<br />

for the foodplant Hermannia jacobifolia (Sterculiaceae) on<br />

which to lay eggs. The female alights on the plant and<br />

immediately begins searching with her antennae for ant<br />

pheromone trails. The ant species has not yet been recorded.<br />

The female will search the 200mm plant from top to bottom and<br />

even the ground around the base. When she is appropriately<br />

stimulated, she usually lays two eggs, side by side, on the stem<br />

<strong>of</strong> the foodplant.<br />

The egg <strong>of</strong> A. d. maseruna is also bun-shaped with a<br />

pronounced network <strong>of</strong> ridges. Initially, the egg is greyishwhite,<br />

becoming purple-brown. The remainder <strong>of</strong> the life history<br />

is unrecorded.<br />

A. d. maseruna is on the wing from November to February<br />

(Henning and Henning 1989).<br />

Threats. The establishment <strong>of</strong> the butterfly reserve at Ruimsig,<br />

specifically for A. d. dentatis, and its presence in the<br />

Suikerbosrand Nature Reserve suggests that this butterfly is no<br />

longer under immediate threat. However, it is essential to<br />

continue monitoring this subspecies, especially as the Ruimsig<br />

population is so small (Henning and Henning 1985, 1989).<br />

Towards this end, over a two week period in December 1989<br />

and January 1990, an estimation was made <strong>of</strong> the population<br />

size at the Ruimsig Reserve. By using a mark-recapture method<br />

it was estimated that the population size was about 400 specimens<br />

on the wing. This was far greater than was expected. This study<br />

will be continued on an annual basis to determine whether the<br />

management methods at the reserve are successful or not.<br />

Recent records for A. d. maseruna from Maseru are sparse.<br />

The most suitable habitats have been used for farming, or<br />

155<br />

washed away by erosion. Another recorded locality is Ladybrand,<br />

apparently along the banks <strong>of</strong> a river where it has not been seen<br />

in recent times. A new locality is at Heilbron, where it has been<br />

found next to a dam in the town and also 10km further north on<br />

a slight slope above a stream. The locality next to the dam is in<br />

the grounds <strong>of</strong> a recently erected old age home, and as this<br />

development continues, the colony may well disappear. The<br />

other locality north <strong>of</strong> Heilbron is apparently safe, as is the<br />

Boons locality, which is in a large, flat, marshy area.<br />

<strong>Conservation</strong>: In 1985 the Roodepoort City Council established<br />

a 12ha reserve at Ruimsig for A. d. dentatis. The butterfly is also<br />

found in the Suikerbosrand Nature Reserve.<br />

No conservation measures are currently in force for A. d.<br />

maseruna.<br />

References<br />

HENNING, S.F. 1983a. Biological groups within the <strong>Lycaenidae</strong> (Lepidoptera).<br />

J. ent. Soc. S. Afr. 46: 65–85.<br />

HENNING, S.F. 1983b. Chemical communication between lycaenid larvae<br />

(Lepidoptera: <strong>Lycaenidae</strong>) and ants (Hymenoptera: Formicidae). J. ent.<br />

Soc. S. Afr. 46: 341–366.<br />

HENNING, S.F. 1984a. The effect <strong>of</strong> ant association on lycaenid larval<br />

duration (Lepidoptera: <strong>Lycaenidae</strong>). Ent. Rec. J. Variation 96: 99–102.<br />

HENNING, S.F. 1984b. Southern African <strong>Butterflies</strong>, with illustrations by<br />

Clare Abbott. Macmillan South Africa, Johannesburg.<br />

HENNING, S.F. 1987. Myrmecophily in lycaenid butterflies (Lepidoptera:<br />

<strong>Lycaenidae</strong>). Ent. Rec. J. Variation 99: 215–222.<br />

HENNING, S.F. and HENNING, G.A. 1985. South Africa's endangered<br />

butterflies. Quagga 10: 16–17.<br />

HENNING, S.F. and HENNING, G.A. 1989. South African Red Data Book–<br />

<strong>Butterflies</strong>. South African National Scientific Programmes Report No.<br />

158, Council for Scientific and Industrial Research, Pretoria, 175 pp.<br />

MURRAY, D.P. 1935. South African <strong>Butterflies</strong>: A Monograph <strong>of</strong> the Family<br />

<strong>Lycaenidae</strong>. John Bale, Sons and Danielsson, London.<br />

PENNINGTON, K.M. 1978. Pennington's <strong>Butterflies</strong> <strong>of</strong> Southern Africa. Ed.<br />

by C.G.C. Dickson, with the collaboration <strong>of</strong> D.M. Kroon. Ad. Donker,<br />

Johannesburg.<br />

TITE, G.E. and DICKSON, C.G.C. 1968a. The Aloeides thyra complex<br />

(Lepidoptera: <strong>Lycaenidae</strong>). Bull. Brit. Mus. (Nat. Hist.)Ent. 21:367–388.<br />

TITE, G.E. and DICKSON, C.G.C. 1968b. The genus Aloeides and allied<br />

genera (Lepidoptera: <strong>Lycaenidae</strong>). Bull. Brit. Mus. (Nat. Hist.) Ent. 29:<br />

225–280.


Erikssonia acraeina Trimen; Subfamily Theclinae, Tribe Aphnaeini<br />

S.F. HENNING 1 , G.A. HENNING 2 and M.J. SAMWAYS 3<br />

1 5 Alexander St., Florida 1709, South Africa<br />

2 17 Sonerend St., Helderdruin 1724, South Africa<br />

3 Department <strong>of</strong> Zoology and Entomology, University <strong>of</strong> Natal, Pietermaritzburg 3200, South Africa<br />

Countries: South Africa, Namibia, Zaire, Zambia.<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable; rare<br />

(Red List).<br />

There are no recent reports <strong>of</strong> known colonies except for the<br />

South African colony, which is thriving on a private farm. This<br />

colony is presently being investigated by the Transvaal Nature<br />

<strong>Conservation</strong> Department, with a view to circumventing the<br />

threat <strong>of</strong> agricultural development destroying the colony. No<br />

conservation measures are being taken to preserve the other<br />

colonies.<br />

Taxonomy and Description: The upperside is orange with a<br />

narrow, black marginal border, a black, subcostal, discal patch<br />

on the forewing and a thin, black, post-discal band on the<br />

hindwing (Pennington 1978). The underside is orange with<br />

thin, post-discal, black lines and scattered black spots. The<br />

female is similar to the male but with a more rounded wing<br />

shape. Forewing lengths: male 15–18mm; female 16–21mm.<br />

Distribution: This species has a scattered distribution, occurring<br />

in Ovamboland in Namibia, and the Waterberg Mountains<br />

(1700m) west <strong>of</strong> Ny lstroom in the western Transvaal. It has also<br />

been recorded from Mongu (Barotse Province, Zambia) and in<br />

Zaire.<br />

Population Size: The Waterberg colony in South Africa is<br />

thought to be strong although no figures on population size are<br />

available. There are no recent reports from the Zambian colony,<br />

and it has not been recorded from Ovamboland, Namibia since<br />

the type was collected.<br />

Habitat and Ecology: E. acraeina flies in open, grassy savannah<br />

with sandy soil, in occasional localities where its foodplant<br />

Gnidia kraussiana (Thymelaeaceae) and host ant (Acantholepis<br />

sp.) occur together. It does not range far from its host plant and<br />

ant. The males establish small territories among the foodplants<br />

in which they can be found throughout most <strong>of</strong> the day. The<br />

females fly randomly within the area, and are as common as the<br />

males. When basking on a sandy patch in the sunshine,<br />

individuals have been seen to lie down on their side maximising<br />

156<br />

exposure. They settle on the ground, on small plants or on grass.<br />

Individuals are <strong>of</strong>ten seen feeding on flowers. E. acraeina flies<br />

slowly and weakly, the bright colour and slow flight indicate<br />

that the species is probably unpalatable through sequestering<br />

toxins from its foodplant. It does not mimic an Acraea as the<br />

name implies but has developed aposematic colouring<br />

independently. The most closely related genus is probably<br />

Aloeides, <strong>of</strong> which most species have tawny-orange colouring.<br />

The eggs are laid in coarse sand at the base <strong>of</strong> the foodplant<br />

near the entrance to the ants' nest. The egg is dome-shaped with<br />

irregular raised convolutions, giving a truffle-like appearance.<br />

Convolutions are absent at the micropyle, which is large, round<br />

and deeply indented. When first laid it is yellowish-ochre, later<br />

darkening to grey or greyish brown.<br />

The larva shelters in the nest during the day, emerging at<br />

night to feed on the foodplant. All instars appear similar. The<br />

body is pinkish-grey with a maroon longitudinal line down the<br />

centre <strong>of</strong> the dorsal surface, flanked on either side by a bluishgreen<br />

area. Laterally, the larvae are marked with regular reddishbrown<br />

markings. The honey gland on the seventh segment is<br />

well developed and the retractile tubercles on the eighth are<br />

white. The sixth (final) instar reaches a maximum length <strong>of</strong><br />

35mm.<br />

Pupation takes place within the ants' nest. The pupa is at<br />

first bright yellow darkening to a deep ochre with a brownish<br />

dorsal line within 48 hours. Pupal length is about 15mm<br />

(Henning 1984).<br />

Adults are on the wing from November to February (Henning<br />

1984; Henning and Henning 1989).<br />

Threats: The Waterberg colony, although strong, is on a<br />

private farm, and therefore may be susceptible to agricultural<br />

development in the long term. There are no recent reports from<br />

the colonies in Zaire or Namibia. Similarly, threats to the<br />

Zairean colony are unknown (Henning and Henning 1989).<br />

<strong>Conservation</strong>: The status <strong>of</strong> this species is currently being<br />

investigated by the Transvaal Nature <strong>Conservation</strong> Department<br />

with assistance from the Lepidopterists' Society <strong>of</strong> Southern<br />

Africa. No conservation measures are being taken to conserve<br />

the species at the other localities.


References<br />

HENNING, S.F. 1984. Life history and behaviour <strong>of</strong> Erikssonia acraeina<br />

Trimen (Lepidoptera: <strong>Lycaenidae</strong>). J. ent. Soc. S. Afr. 47: 337–342.<br />

HENNING, S.F. and HENNING, G.A. 1989. South African Red Data Book-<br />

157<br />

<strong>Butterflies</strong>. South African National Scientific Programmes Report No.<br />

158, Council for Scientific and Industrial Research, Pretoria, 175 pp.<br />

PENNINGTON, K.M. 1978. Pennington's <strong>Butterflies</strong> <strong>of</strong> Southern Africa. Ed.<br />

by C.G.C. Dickson, with the collaboration <strong>of</strong> D.M. Kroon. Ad. Donker,<br />

Johannesburg.


Alaena margaritacea Eltringham; Subfamily Lipteninae, Tribe<br />

Pentilini<br />

Country: South Africa.<br />

S.F. HENNING 1 , G.A. HENNING 2 and M.J. SAMWAYS 3<br />

1 5 Alexander St., Florida 1709, South Africa<br />

2 17 Sonerend St., Helderdruin 1724, South Africa<br />

3 Department <strong>of</strong> Zoology and Entomology, University <strong>of</strong> Natal, Pietermaritzburg 3200, South Africa<br />

Status and <strong>Conservation</strong> Interest: Status – vulnerable (Red<br />

List).<br />

This liptenine is endemic to one locality in the northeastern<br />

Transvaal. Two colonies are known high in the Strydpoort<br />

Mountains where the butterfly inhabits the steep grassy slopes<br />

and lichen-covered rocks 400m from the summit. Planting <strong>of</strong><br />

stands <strong>of</strong> pine trees in the area poses a threat, although the<br />

Transvaal Nature <strong>Conservation</strong> Department is aware <strong>of</strong> the<br />

situation.<br />

Taxonomy and Description: This is a small species (Clark and<br />

Dickson 1971; Murray 1935; Pennington 1978) with elongated<br />

forewings. The upperside is black with a broad orange band,<br />

which is very broad in the female, almost reaching the base. The<br />

underside <strong>of</strong> the hindwing is creamy white with an intricate,<br />

lace-like pattern <strong>of</strong> thin black lines. Forewing lengths: male<br />

12–13.5mm; female 14–15mm.<br />

Distribution: A. margaritacea is endemic to South Africa, and<br />

is known only from one locality near Haenertsburg in the<br />

northeastern Transvaal (Henning and Henning 1989).<br />

Population Size: Although the two colonies near Haenertsburg<br />

only cover an area <strong>of</strong> about lha, there can be numerous<br />

individuals.<br />

Habitat and Ecology: Two secluded colonies are known from<br />

the slopes <strong>of</strong> the Strydpoort Mountains about 400m below the<br />

peaks. A. margaritacea flies on the steep grassy slopes with<br />

large lichen-covered rocks (Swanepoel 1953). Near midday,<br />

the males have been recorded ascending almost to the mountain<br />

tops, where they establish small territories at the base <strong>of</strong> rocky<br />

ridges just below the peaks. A. margaritacea has a weak<br />

fluttering flight, and when disturbed, it flies a few metres before<br />

settling again on a grass stem. If repeatedly disturbed, it may fly<br />

for a hundred metres before settling again. The males normally<br />

158<br />

establish territories in the breeding area, perching on grass<br />

stalks and fluttering around the grass. When another male<br />

enters its territory, it sometimes chases the intruder away. The<br />

female flutters slowly throughout the breeding area, searching<br />

for suitable lichen on which to lay her eggs.<br />

The foodplant is probably rock lichen, although the female<br />

sometimes lays on rocks and stones supporting only a little<br />

lichen. The eggs are laid singly or in small clusters. The eggs are<br />

0.9mm in diameter, 0.4mm high, and purple-brown. There are<br />

four rings <strong>of</strong> 14, round indentations on each egg. Those at the<br />

micropyle are narrow and elongated. Nothing is known <strong>of</strong><br />

larval behaviour. The first instar is purple-brown with a pale<br />

yellow neck-shield and white humps which bear the outer<br />

dorsal and lateral setae. The head is purple-brown. The<br />

subsequent instars and pupa are unrecorded.<br />

The adult is on the wing in December and January, with a<br />

peak towards the end <strong>of</strong> December.<br />

Threats: The planting <strong>of</strong> plantation pine trees is a serious threat<br />

to this species.<br />

<strong>Conservation</strong>: The Transvaal Nature <strong>Conservation</strong><br />

Department is aware <strong>of</strong> the threats to this species, and is<br />

currently monitoring it.<br />

References<br />

CLARK, G.C. and DICKSON, C.G.C. 1971. Life Histories <strong>of</strong> the South<br />

African Lycaenid <strong>Butterflies</strong>. Purnell and Sons, Cape Town.<br />

HENNING, S.F. and HENNING, G.A. 1989. South African Red Data Book–<br />

<strong>Butterflies</strong>. South African National Scientific Programmes Report No.<br />

158, Council for Scientific and Industrial Research, Pretoria, 175 pp.<br />

MURRAY, D.P. 1935. South African <strong>Butterflies</strong>: A Monograph <strong>of</strong> the Family<br />

<strong>Lycaenidae</strong>. John Bale, Sons and Danielsson, London.<br />

PENNINGTON, K.M. 1978. Pennington's <strong>Butterflies</strong> <strong>of</strong> Southern Africa. Ed.<br />

by C.G.C. Dickson, with the collaboration <strong>of</strong> D.M. Kroon. Ad. Donker,<br />

Johannesburg.<br />

SWANEPOEL,D.A. 1953. <strong>Butterflies</strong> <strong>of</strong> South Africa: Where, When and How<br />

they Fly. Maskew Miller, Cape Town.


Orachrysops (Lepidochrysops) ariadne (Butler); Subfamily<br />

Polyommatinae, Tribe Polyommatini<br />

Country: South Africa.<br />

S.F. HENNING 1 , G.A. HENNING 2 and M.J. SAMWAYS 3<br />

' 5 Alexander St., Florida 1709, South Africa<br />

2 17 Sonerend St., Helderdruin 1724, South Africa<br />

3 Department <strong>of</strong> Zoology and Entomology, University <strong>of</strong> Natal, Pietermaritzburg 3200, South Africa<br />

Status and <strong>Conservation</strong> Interest: Status – rare; endangered<br />

(Red List).<br />

O. ariadne is a highly localised Natal endemic. One <strong>of</strong> the<br />

two recorded colonies has become extinct, and the other occupies<br />

1 hectare on private land. Neglectful management <strong>of</strong> the site is<br />

posing a serious threat. Much more biological information on<br />

the species is required, as is a management plan for its protection.<br />

Taxonomy and Description: This species has rounded wings.<br />

The males are dull violaceous-blue on the upperside with<br />

narrow dark brown margins. The underside is dark greyishbrown<br />

with black spots and a distinct line <strong>of</strong> clearly-defined<br />

white post-discal marks. The female is brown on the upperside<br />

with blue areas reduced to the basal half <strong>of</strong> the wings; the<br />

underside is similar to that <strong>of</strong> the male. Forewing lengths: male<br />

13–17 mm; female 13–19 mm.<br />

Distribution: O. ariadne is endemic to Natal (Henning and<br />

Henning 1989). It has only been found in the Karklo<strong>of</strong> District,<br />

although previously it was recorded nearby at Balgowan. A few<br />

specimens which could represent this species have been recorded<br />

from Nkandla near Eshowe.<br />

Population Size: The population size <strong>of</strong> the only confirmed<br />

colony (at Karklo<strong>of</strong> Falls) is not known.<br />

Habitat and Ecology: O. ariadne inhabits a one hectare area <strong>of</strong><br />

steep grassy slopes adjacent to forests (Pennington 1978). It<br />

occurs in tall grass on the north side <strong>of</strong> the stream running down<br />

to the top <strong>of</strong> the Karklo<strong>of</strong> Falls. On the south-facing steep slope,<br />

among the tall Hyparrhenia spp. grasses, the foodplant<br />

Indig<strong>of</strong>era astragalina (Fabaceae) is found (Swanepoel 1953).<br />

159<br />

O. ariadne is a fast, low flier, but may fly to a height <strong>of</strong> two<br />

metres to clear the tall grassheads. The males patrol up and<br />

down the valley, dodging in and out <strong>of</strong> the tall grass, sometimes<br />

venturing into the adjacent valley across the stream, but always<br />

returning to where the foodplant grows. They do not appear to<br />

be strongly territorial.<br />

This species is on the wing from l000h to 1430h. The<br />

females spend much time looking for the foodplant on which to<br />

lay their eggs. They fly more slowly than the males, and are<br />

always found in the vicinity. Little is known <strong>of</strong> the life-history,<br />

although early instars have been seen to feed on the foodplant<br />

before going down, as third instars, into an unidentified ant nest<br />

to feed on the brood.<br />

Threats: The habitat is being overgrown. In the past, vertebrate<br />

herbivores and natural fires kept the site in a more open, earlier<br />

successional stage suitable for the foodplant and ant host.<br />

<strong>Conservation</strong>: Careful monitoring <strong>of</strong> the site and active<br />

management is advisable to prevent the colony disappearing.<br />

No formal conservation measures are currently in force, aside<br />

from some interest by the owners <strong>of</strong> the land.<br />

References<br />

HENNING, S.F and HENNING, G.A. 1989. South African Red Data Book -<br />

<strong>Butterflies</strong>. South African National Scientific Programmes Report No.<br />

158, Council for Scientific and Industrial Research, Pretoria. 175 pp.<br />

PENNINGTON, K.M. 1978. Pennington's <strong>Butterflies</strong> <strong>of</strong> Southern Africa. Ed.<br />

by C.G.C. Dickson, with the collaboration <strong>of</strong> D.M. Kroon. Ad. Donker,<br />

Johannesburg.<br />

SWANEPOEL,D.A. 1953. <strong>Butterflies</strong> <strong>of</strong> South Africa: Where, When and How<br />

they Fly. Maskew Miller, Cape Town.


Country: Australia<br />

Hypochrysops C. and R. Felder<br />

D.P.A. SANDS<br />

Division <strong>of</strong> Entomology, CSIRO, Meiers Road, Indooroopilly, Queensland 4068, Australia<br />

Status and <strong>Conservation</strong> Interest: Status – several species<br />

endangered or rare.<br />

Habitat destruction has affected seven species whose survival<br />

is now considered to be threatened in four Australian States<br />

(Table 1, last column). The most threatened species in the<br />

genus, H. piceatus, is only known from two localities in<br />

southern Queensland, one formerly at Millmerran but now<br />

cleared for farming and the other consisting <strong>of</strong> a few kilometres<br />

<strong>of</strong> roadside savannah near Leyburn. Together with Paralucia<br />

spinifera Edwards & Common, these two butterflies are the<br />

most 'at risk' <strong>of</strong> all the Australian butterflies. It is possible that<br />

other localities will be found for H. piceatus but none has yet<br />

been discovered despite searches <strong>of</strong> suitable intact open forest<br />

containing the foodplant, Casuarina luehmannii. Fortunately,<br />

the butterflies are not uncommon in some years and have<br />

maintained their abundance despite heavy collecting at times<br />

by amateurs.<br />

Taxonomy and Description: Species <strong>of</strong> Hypochrysops are<br />

renowned among lepidopterists for the distinctive and beautiful<br />

patterns on the undersides <strong>of</strong> their wings. Unlike the closely<br />

related genus Philiris Rober, most species <strong>of</strong> Hypochrysops are<br />

easily distinguished and were described before the turn <strong>of</strong> the<br />

century. All except one (H. piceatus Kerr, Macqueen & Sands)<br />

<strong>of</strong> the species occurring in Australia were described by the time<br />

Waterhouse's 'WhatButterfly is That?' (1932) was published.<br />

However, the specific and subspecific status <strong>of</strong> some populations<br />

has been reviewed recently (Sands 1986) and some taxonomic<br />

identities changed.<br />

Distribution: The majority <strong>of</strong> the 57 species recognised (Sands<br />

1986), occur on the neighbouring islands north <strong>of</strong> Australia and<br />

only one (H. coelisparsus [Butler]) occurs west <strong>of</strong> Wallace's<br />

line. Six <strong>of</strong> the 18 species <strong>of</strong> Hypochrysops in Australia are<br />

endemic while the remainder are also found on mainland New<br />

Guinea.<br />

Population Size: Not known.<br />

160<br />

Habitat and Ecology: The life history <strong>of</strong> H. ignitus ignitus<br />

(Leach) was described by Waterhouse (1932) and it differs only<br />

slightly for subspecies erythrinus. Adults oviposit on small (up<br />

to 2m) plants <strong>of</strong> Eucalyptus confertiflora or Acacia spp. and<br />

their larvae shelter by day in byres <strong>of</strong> the attendant ant,<br />

Iridomyrmex nitidus, constructed on the stems and trunk <strong>of</strong> the<br />

foodplant. At the end <strong>of</strong> the wet season populations normally<br />

stabilise after the dry season stress – a time when much <strong>of</strong> their<br />

habitat is now burnt intentionally as a precaution against wild<br />

fires.<br />

Threats: Not all species 'hilltop' but for those that do, human<br />

activities such as clearing and installation for radio and television<br />

equipment, forestry observation towers and mining <strong>of</strong> rock<br />

outcrops <strong>of</strong>ten result in destruction <strong>of</strong> the hilltop habitats and<br />

local extinctions. The most susceptible to these activities are the<br />

subspecies <strong>of</strong> H. delicia Hewitson and H. ignitus (Leach).<br />

It is likely that the intentional annual burning <strong>of</strong> savannah in<br />

the Northern Territory near Darwin contributes to the extreme<br />

rarity <strong>of</strong> H, ignitus erythrinus (Waterhouse & Lyell). The firing<br />

is carried out by conservation authorities in the belief that the<br />

original inhabitants did so as a method for hunting wildlife.<br />

However, these fires now started at the end <strong>of</strong> every wet season<br />

are quite devastating to several insect species and particularly<br />

to H. ignitus erythrinus.<br />

At the end <strong>of</strong> the wet season populations <strong>of</strong> H, ignitus ignitus<br />

normally stabilise after the dry season stress. The abundance <strong>of</strong><br />

this subspecies, dependant on suckers and low regrowth <strong>of</strong> the<br />

larval foodplants, has been dramatically reduced since routine<br />

fuel reduction burning began. It is said that vertebrates are not<br />

affected by these 'slow burns' since they are sufficiently mobile<br />

to escape from the advancing flames. Immature stages <strong>of</strong><br />

arthropods, on the other hand, are incinerated when surrounded<br />

by combustible materials close to the ground. There is an urgent<br />

need to review these burn policies and to take account <strong>of</strong> sessile<br />

organisms, perhaps by providing fire-free areas where dry<br />

country habitats <strong>of</strong> rare insect fauna are allowed to stabilise.<br />

The Australian species <strong>of</strong> Hypochrysops are local,<br />

uncommon or rare, occurring in undisturbed habitats and very<br />

few if any are able to adapt to encroaching urbanisation. H.


pythias (Felder & Felder) occurs at times in hundreds in Papua<br />

New Guinea, especially where its foodplant Commersonia<br />

bartramia has regrown after human disturbance, but the<br />

Australian subspecies euclides has never been observed in such<br />

numbers even though the food plant is the same and <strong>of</strong>ten a<br />

predominant regrowth plant. It is an unusual species as adults<br />

are known to aggregate in large numbers at night in Papua New<br />

Guinea.<br />

For those species that congregate on undisturbed hilltops<br />

males are sometimes seen in numbers but the females are<br />

always uncommon. Hilltops are selected as locations where<br />

unmated female adults can, with a degree <strong>of</strong> certainty, find a<br />

mate among the competing males.<br />

Australian tropical and subtropical rainforests have been<br />

depleted to a fraction <strong>of</strong> their original area, and conservation <strong>of</strong><br />

the remainder is essential if the survival <strong>of</strong> several species <strong>of</strong><br />

Hypochrysops is not to be threatened. The fringing rainforests<br />

along the Rocky River have suffered from gold mining activities<br />

placing at risk survival <strong>of</strong> the unique H. theon cretatus Sands.<br />

Nowhere else is this recently-described subspecies known to<br />

occur. It is the most southern population <strong>of</strong> a species found<br />

throughout mainland New Guinea and is the most distinctive <strong>of</strong><br />

all <strong>of</strong> the seven subspecies recognised (Sands 1986). Fortunately,<br />

much <strong>of</strong> the habitat for another subspecies, H. theon medocus<br />

(Fruhstorfer), and the related H. hippuris nebulosis Sands is not<br />

at risk. However, the rainforest margin species, H. miskini<br />

miskini (Waterhouse) is seriously threatened by habitat<br />

destruction in southeastern Queensland. Formerly common<br />

near Burleigh Heads, on the Coomera River and near Rainbow<br />

Beach, the species has disappeared from most localities and is<br />

becoming rare at the remainder. Although slightly different in<br />

morphology from the southern populations, populations <strong>of</strong> H.<br />

miskini from the northern localities are not so threatened.<br />

At the edge <strong>of</strong> its range in northern NSW, H. digglesii<br />

(Hewitson) is now confined mainly to moist savannah near<br />

Broken Head. Unfortunately, although most <strong>of</strong> the nearby<br />

rainforest is included in a National Park, the breeding sites for<br />

this species are at present threatened by local development<br />

proposals.<br />

It is important to note that this is the only locality remaining<br />

intact for this species in NSW and that it differs quite considerably<br />

in appearance from the more abundant populations in<br />

Queensland.<br />

Coastal paperbark (Melaleuca spp.) swamps are habitats for<br />

several species destroyed on a large scale in recent years. H.<br />

apollo apollo M iskin was once quite abundant at a few paperbark<br />

localities including Cardwell, the southern edge <strong>of</strong> its range and<br />

at Port Douglas, northern Queensland. However, forestry<br />

activities, mainly clearing and planting with Pinus sp. and earth<br />

works for a canal development at the first locality and clearing<br />

for a golflinks at the second, have seriously damaged the major<br />

breeding sites for this unusual butterfly. H. apollo apollo has<br />

suffered from clearing <strong>of</strong> the paperbark swamps that support<br />

the epiphytic plant, Myrmecodia beccari, the larval foodplant.<br />

Table 1. Hypochrysops spp. currently known to survive at five or fewer localities in States where they were formerly widely distributed.<br />

Taxon<br />

H. apelles apelles<br />

H. apollo apollo<br />

H. byzos byzos<br />

H. cleon<br />

H. digglesii<br />

H. epicurus<br />

H. hippuris nebulosus<br />

H. ignitus ignitus<br />

H. ignitus erythrinus<br />

H. piceatus<br />

H. theon cretatus<br />

Key<br />

NSW<br />

Q<br />

Q<br />

Q<br />

NSW<br />

NSW<br />

Q<br />

SA<br />

NT<br />

Q<br />

Q<br />

State† with nos<br />

localities<br />

intact<br />

† NSW New South Wales * MG mangroves<br />

0 Queensland SV savannah<br />

NT Northern Territory RF rainforest<br />

SA South Australia<br />

V Victoria<br />

WA Western Australia<br />

c. 4<br />

c. 5<br />

c.3<br />

1<br />

1<br />

c.5<br />

1<br />

1<br />

2<br />

1<br />

1<br />

161<br />

Present<br />

elsewhere<br />

(state*)<br />

Q, NT<br />

–<br />

NSW<br />

–<br />

Q<br />

Q<br />

–<br />

Q, V, NSW<br />

WA<br />

–<br />

–<br />

Type*<br />

MG<br />

SV<br />

SV<br />

RF<br />

SV<br />

MG<br />

RF<br />

SV<br />

SV<br />

SV<br />

RF<br />

Habitats<br />

At risk<br />

+<br />

+<br />

–<br />

–<br />

+<br />

+<br />

–<br />

+<br />

+<br />

+<br />

+


Another factor that is decreasing the abundance <strong>of</strong> the butterfly<br />

is removal <strong>of</strong> the bulbous foodplants by collectors and who cut<br />

them up in search <strong>of</strong> pupae. Colonisation <strong>of</strong> the bulbs by the<br />

exotic ant Pheidole megacephala may also affect survival <strong>of</strong> H.<br />

apollo since it displaces the native Iridomyrmex cordatus, the<br />

natural bulb inhabitant. Indeed, P. megacephala may eventually<br />

threaten the survival <strong>of</strong> ant plants, as it is known to destroy the<br />

developing seeds. This is probably one <strong>of</strong> the very few examples<br />

<strong>of</strong> a conservation issue for Australian butterflies involving<br />

threat to the breeding sites by collectors. The same habitat is<br />

shared with the much more abundant H. narcissus narcissus<br />

(F.) which breeds on mistletoes, a species also present in and<br />

breeding on mangroves in northern Queensland.<br />

Mangroves are important breeding sites for a number <strong>of</strong><br />

interesting lycaenid butterflies. In southeastern Queensland<br />

and northern NSW, the coppery H. apelles apelles (F.) only<br />

occurs near saltwater swamps where the larvae feed on several<br />

species <strong>of</strong> mangroves. The species was once very common near<br />

Southport and on the Tweed River but its numbers have declined<br />

162<br />

in recent years and it has disappeared from several localities.<br />

Housing developments, clearing <strong>of</strong> mangroves and possibly,<br />

spraying with insecticides have resulted in the species becoming<br />

scarce in NSW and uncommon in southern Queensland. Further<br />

north in Queensland a wide range <strong>of</strong> plants are food for larvae<br />

and the species is not at risk. Another species, H. epicurus<br />

Miskin, has suffered from similar destruction <strong>of</strong> its only<br />

mangrove (Avicennia marina) habitat in southeastern<br />

Queensland and NSW. This species has become quite rare at<br />

localities where it was formerly very common due to habitat<br />

interference.<br />

References<br />

SANDS, D.P.A. 1986. A revision <strong>of</strong> the genus Hypochrysops C. & R. Felder<br />

(Lepidoplera: <strong>Lycaenidae</strong>). Entomonograph No. 7. E.J. Brill, Leiden.<br />

WATERHOUSE, G.A. 1932. What Butterfly is That? Angus & Robertson,<br />

Sydney.


Illidge's Ant-Blue, Acrodipsas illidgei (Waterhouse and Lyell)<br />

Country: Australia.<br />

P.R. SAMSON<br />

Bureau <strong>of</strong> Sugar Experiment Stations, P.O. Box 651, Bundaberg, Queensland 4650, Australia<br />

Status and <strong>Conservation</strong> Interest: Status – rare (Common<br />

and Waterhouse 1981).<br />

Its life cycle is unusual, larvae being obligate<br />

myrmecophages. It has a restricted distribution and occurs<br />

almost exclusively among mangroves, a habitat which is under<br />

threat from development in the populous area where the species<br />

is known. A proposal for a canal estate and residential subdivision<br />

was rejected in a local government court decision in Brisbane<br />

in 1989, based in part on the need to preserve colonies <strong>of</strong> A.<br />

illidgei occurring in the area. In July 1990, A. illidgei became<br />

the first butterfly to be designated as 'Permanently Protected<br />

Fauna' in Queensland.<br />

Taxonomy and Description: A. illidgei was originally described<br />

as a subspecies <strong>of</strong> A. myrmecophylla (Waterhouse & Lyell),<br />

from which it differs by its larger size, the broader markings<br />

beneath the wings and the male genitalia (Kerr, Macqueen and<br />

Sands 1968).<br />

Acrodipsas Sands contains seven described species, all<br />

endemic to Australia (Common and Waterhouse 1981). They<br />

were previously referred to as Pseudodipsas C. and R. Felder.<br />

Four species have been described since 1965. Two more probable<br />

species await description (D. Sands, pers. comm.). All species<br />

except A. illidgei are known mostly from males collected on<br />

hilltops in open Eucalyptus forest. The life histories are poorly<br />

understood, and only A. illidgei and two others have been reared<br />

from the immature stages.<br />

Distribution: A. illidgei is restricted to coastal areas, mostly in<br />

mangrove habitats. For many years the species had only been<br />

recorded between Brisbane and Burleigh Heads in southern<br />

Queensland, but there are recent records from Mary River Heads<br />

near Maryborough, Queensland (Manskie and Manskie 1989)<br />

and Brunswick Heads, New South Wales (G. Miller pers. comm.).<br />

Population Size: Almost all recent specimens have been<br />

collected at Burleigh Heads or at Redland Bay near Brisbane.<br />

There are no estimates <strong>of</strong> the population sizes at these sites.<br />

Newly-emerged female <strong>of</strong> Acrodipsas illidgei (photo by P.R. Samson). Final instar larva <strong>of</strong> Acrodipsas illidgei with associated ants and brood<br />

(photo by P.R. Samson).<br />

163


Habitat and Ecology: The colonies at Burleigh and Redland<br />

Bay occur in communities dominated by the grey mangrove<br />

Avicennia marina. Other records from Mary River Heads,<br />

Hay's Inlet near Brisbane, and Brunswick Heads are also from<br />

mangrove areas.<br />

Adults <strong>of</strong> A. illidgei are small and inconspicuous, and are<br />

not easy to see among the dense mangrove vegetation. My<br />

observations at Redland Bay suggest that breeding occurs in<br />

patches <strong>of</strong> trees. I have found 25 eggs laid on a single mangrove<br />

tree with smaller numbers on adjacent trees. On that occasion<br />

the vast majority <strong>of</strong> trees apparently carried no eggs, despite the<br />

presence <strong>of</strong> ant colonies in some <strong>of</strong> them. Whether these<br />

patches remain stable over time is unknown.<br />

Of the immature stages, only the eggs are in exposed<br />

positions, and they are difficult to locate. Eggs are laid on<br />

branches or under loose bark <strong>of</strong> trees colonised by the associated<br />

ants. Eggs are aggregated on the particular trees used for<br />

breeding. They hatch in about one week during summer.<br />

The first instar larvae are carried into ant nests by the small<br />

ant, Crematogaster sp. (laeviceps group) (Samson 1987,1989).<br />

The nests occur in cavities (mostly in borer holes) in the wood<br />

<strong>of</strong> living trees. There, the larvae prey on the ant brood throughout<br />

their development. Pupae also occur inside the nest and several<br />

larvae or pupae may be found together in a hollow branch.<br />

Larvae and pupae occurring in these nests can only be found by<br />

destroying the branches and ant colonies in which they live.<br />

Almost all have been found inside nests in A. marina. However,<br />

several have been found beneath bark <strong>of</strong> Eucalyptus (Smales<br />

and Ledward 1942), indicating that there is no special<br />

relationship with Avicennia.<br />

Collecting records suggest that there are at least two<br />

generations each year, with maximum frequency <strong>of</strong> capture in<br />

September and December to February: no adults have been<br />

taken during the colder months <strong>of</strong> May through July.<br />

Adults have been seen feeding at flowers (Hagan 1980).<br />

Like some other species <strong>of</strong> Acrodipsas, females possess many<br />

fully developed eggs at emergence from the pupa (Sands 1979).<br />

Another lycaenid, Hypochrysops apelles (Fabricius), has an<br />

obligate relationship with Crematogaster sp. (laeviceps group)<br />

in southern Queensland. That butterfly occurs continuously<br />

along the coast to northern Queensland (Common and<br />

Waterhouse 1981), probably attended by the same ant. It is<br />

surprising, therefore, that A. illidgei has such a restricted<br />

distribution.<br />

Threats: Survival <strong>of</strong> A. illidgei is threatened by the loss <strong>of</strong><br />

mangroves to residential development. The species' distribution<br />

coincides with the most populous region <strong>of</strong> Queensland, where<br />

coastal development is proceeding at a rapid rate. Fogging with<br />

insecticides to control biting insects that breed in mangroves<br />

may also threaten A. illidgei.<br />

Collecting poses little threat to the species. Only a small<br />

number <strong>of</strong> collectors have visited the breeding sites. The<br />

butterflies do not have a sharply defined flight period and are<br />

very difficult to find and capture in the habitat. Collecting <strong>of</strong><br />

immature stages destroys the ant galleries that are searched.<br />

164<br />

However, it is only the galleries in the smaller, dead branches<br />

<strong>of</strong> the mangroves that are readily opened up by collectors: the<br />

greater part <strong>of</strong> the ant nests in the large branches and trunks <strong>of</strong><br />

the mangroves are inaccessible.<br />

<strong>Conservation</strong>: Part <strong>of</strong> the habitat <strong>of</strong> A. illidgei at Burleigh<br />

Heads is protected within an environmental park established to<br />

preserve what little remains <strong>of</strong> mangroves in the area. Whether<br />

A. illidgei still breeds within the park is not known.<br />

Recently, approval for a canal estate proposal at Point<br />

Halloran, Redland Bay was rejected because <strong>of</strong> deleterious<br />

effects on the environment. The proposed development was to<br />

have included 610 residential allotments <strong>of</strong> which 385 were to<br />

have canal frontages on 106ha. The presence <strong>of</strong> A. illidgei in the<br />

area was one factor in the court's decision to reject the proposal<br />

(The Courier Mail, 17 June 1989). A small part <strong>of</strong> this area has<br />

since been rezoned for residential development, but the<br />

remainder is likely to become an environmental park.<br />

The colony at Redland Bay that is most <strong>of</strong>ten visited by<br />

lepidopterists is near to, but not contained within, the<br />

development proposal. There is at present no habitat protection<br />

afforded to this site.<br />

The mangrove habitat <strong>of</strong> A. illidgei at Mary River Heads is<br />

part <strong>of</strong> an area nominated for world heritage listing. A request<br />

from local residents for insecticidal fogging to control biting<br />

midges has been denied because <strong>of</strong> possible harm to the<br />

butterflies (The Courier Mail, 2 May 1992).<br />

The designation <strong>of</strong> A. illidgei as 'Permanently Protected<br />

Fauna' under the Queensland Fauna <strong>Conservation</strong> Act 1974–79<br />

on July 21, 1990 gives it an unusually high and controversial<br />

status (Monteith 1990). This status has otherwise been accorded<br />

to a few high-pr<strong>of</strong>ile vertebrates, such as the koala and platypus.<br />

Permits are needed to keep specimens in private collections,<br />

and a separate permit is required in order to move specimens to<br />

different premises, or interstate. Permits may be obtained for<br />

scientific study <strong>of</strong> the species but a fear is that the degree <strong>of</strong><br />

formality imposed is likely to deter the interest <strong>of</strong> enthusiastic<br />

amateurs who have contributed so much to knowledge <strong>of</strong><br />

Australian <strong>Lycaenidae</strong>.<br />

References<br />

COMMON, I.F.B. and WATERHOUSE, D.F. 1981. <strong>Butterflies</strong> <strong>of</strong> Australia.<br />

(2nd edn.) Angus and Robertson, Sydney.<br />

HAGAN, C.E. 1980. Recent records <strong>of</strong> Acrodipsas illidgei (Waterhouse and<br />

Lyell) (Lepidoptera: <strong>Lycaenidae</strong>) from the Brisbane area, Queensland.<br />

Aust. ent. Mag. 7: 39.<br />

KERR, J.F.R., MACQUEEN, J. and SANDS, D.P. 1968. The specific status<br />

<strong>of</strong> Pseudodipsas illidgei Waterhouse and Lyell stat. n. (Lepidoptera:<br />

<strong>Lycaenidae</strong>). J. Aust. ent. Soc. 7: 28.<br />

MANSKIE, R.C. and MANSKIE, N. 1989. New distribution records for four<br />

Queensland <strong>Lycaenidae</strong> (Lepidoptera). Aust. ent. Mag. 16: 98.<br />

MONTEITH, G. 1990. Another butterfly protected in Queensland. Myrmecia<br />

August 1990: 153–154.<br />

SAMSON, P.R. 1987. The blue connection: butterflies, ants and mangroves.<br />

Aust. nat. Hist. 22: 177–181.


SAMSON, P.R. 1989. Morphology and biology <strong>of</strong> Acrodipsas illidgei<br />

(Waterhouse and Lyell), a myrmecophagous lycaenid (Lepidoptera:<br />

<strong>Lycaenidae</strong>: Theclinae). J. Aust. ent. Soc. 28: 161–168.<br />

SANDS, D.P.A. 1979. A new genus, Acrodipsas, for a group <strong>of</strong> <strong>Lycaenidae</strong><br />

(Lepidoptera) previously referred to Pseudodipsas C. & R. Felder, with<br />

165<br />

descriptions <strong>of</strong> two new species from northern Queensland. J. Aust. ent.<br />

Soc. 18: 251–265.<br />

SMALES, M. and LEDWARD, C.P. 1942. Notes on the life-histories <strong>of</strong> some<br />

lycaenid butterflies – Part I. Qd. Nat. 12: 14–18.


Country: Australia.<br />

The Elthani Copper, Paralucia pyrodiscus lucida Crosby<br />

T.R. NEW<br />

Department <strong>of</strong> Zoology, La Trobe University, Bundoora, Victoria 3083, Australia<br />

Status and <strong>Conservation</strong> Interest: Status – rare, vulnerable.<br />

This local subspecies was feared to be extinct in the<br />

Melbourne area <strong>of</strong> southeastern Australia before a thriving<br />

colony was found in January 1987 on land already subdivided<br />

for imminent housing development. A major campaign was<br />

undertaken to (a) acquire and reserve the habitat, (b) search for<br />

other colonies and (c) design a management plan for the<br />

butterfly. This case broke new ground in increasing invertebrate<br />

conservation awareness in Victoria.<br />

Taxonomy and Description: The subspecies was described<br />

(Crosby 1951) from the Eltham (37°43'S, 149°09'E) /<br />

Greensborough area <strong>of</strong> outer northeastern Melbourne, and<br />

differs in the amount <strong>of</strong> copper scaling from the nominate<br />

subspecies, P. p. pyrodiscus (Doubleday) which occurs in parts<br />

<strong>of</strong> eastern Australia as far north as southern Queensland.<br />

Colonies around Kiata (36°22'S, 141°48'E) and Castlemaine<br />

(37°04'S, 144°13'E) are also at present referred to this<br />

subspecies. The extra-Victorian limits <strong>of</strong> the subspecies are not<br />

wholly clear: at present no other populations are referred to<br />

lucida formally.<br />

Paralucia Waterhouse and Turner contains three species,<br />

all endemic to Australia. One, P. aurifera (Blanchard), is<br />

widespread but local and the other two are rare.<br />

Distribution: The subspecies is very restricted in isolated parts<br />

<strong>of</strong> Victoria. Eight discrete colonies are known around Eltham;<br />

the subspecies otherwise occurs only at Kiata (six colonies<br />

within 3km <strong>of</strong> each other; one at nearby Salisbury) and<br />

Castlemaine (a single colony only) (Braby et al. 1992).<br />

Population Size: P. p. lucida was discovered near Melbourne<br />

in 1938. Accumulated collector wisdom shows that it was taken<br />

more or less regularly over the following decade or so but<br />

declined markedly from about 1950 onwards and thereafter<br />

became extremely rare. It was believed to be extinct in recent<br />

years until the discovery <strong>of</strong> a thriving colony near Melbourne<br />

in 1987.<br />

166<br />

All presently known colonies are very isolated, and the<br />

population structure is essentially closed. There is very little<br />

possibility <strong>of</strong> interchange <strong>of</strong> adults between most nearby<br />

suburban colonies, and none over the broader scale <strong>of</strong><br />

distribution. Most <strong>of</strong> the Eltham colonies are small and seem<br />

unlikely to be viable in the long term. Population sizes have<br />

been estimated both by counts <strong>of</strong> adult butterflies during the<br />

flight season and nocturnal counts <strong>of</strong> feeding caterpillars<br />

(Vaughan 1987, 1988) and as the largest (presumed viable)<br />

colonies occupy only a few hundred square metres, these counts<br />

are likely to be reasonably accurate. The major Eltham colony,<br />

on the subdivision land, contained an estimated 300–500 larvae.<br />

At the other extreme only six butterflies were observed in a<br />

colony on private land. Several intermediate sized colonies<br />

each had populations estimated at 100–150 individuals. In<br />

contrast, the Kiata colonies contained 'many hundreds' <strong>of</strong><br />

individuals, and there are 100–200 at Castlemaine. In 1988, the<br />

State population <strong>of</strong> the Eltham Copper was about 2600<br />

individuals (Braby and Crosby in prep.).<br />

Habitat and Ecology: Some colonies are confined to open<br />

forest areas, generally on north to west-facing slopes, and<br />

others occur on highly degraded areas such as roadsides. The<br />

life cycle appears to be limited obligatorily to a spiny dwarfed<br />

form <strong>of</strong> Sweet Bursaria (Bursaria spinosa: Pittosporaceae) as<br />

the sole larval foodplant, and to plants associated with nest<br />

chambers <strong>of</strong> the ant genus Notoncus. There is one major<br />

generation each year, with adults present from late November<br />

to early February, and a possible small, partial second generation<br />

represented by a few adults in March (Braby 1990).<br />

Eggs are laid singly or in small groups on or near the larval<br />

foodplant. Caterpillars hatch after about two weeks and are<br />

nocturnal feeders: they retreat to Notoncus chambers at the base<br />

<strong>of</strong> the plants during the daytime and are regularly tended by the<br />

ants, these being N. enormis Szabo at Eltham but N.<br />

ectatommoides (Forel) at Kiata. Caterpillars pupate in the ant<br />

chambers after five instars, and the pupal stage <strong>of</strong> the major<br />

generation lasts about a month.<br />

Adult P. p. lucida take nectar from various flowers, including<br />

Bursaria. They seem to be aggressive to other butterflies, and<br />

males actively pursue females.


Both the stunted Bursaria and Notoncus ants are widespread<br />

in Victoria, and there is seemingly no shortage <strong>of</strong> sites suitable<br />

for P. p. lucida. Its restricted distribution is at present difficult<br />

to explain.<br />

Threats: Recent declines appear to be attributable solely to<br />

urbanisation, with associated removal <strong>of</strong> native vegetation,<br />

fragmentation and destruction. The habitats <strong>of</strong> some former<br />

colonies near Eltham are now housing estates and this was<br />

clearly the major threat to the newly discovered colonies. The<br />

largest colony was reduced substantially during 1988–89. The<br />

very small colony noted above was in a suburban garden, but<br />

many such gardens are rapidly converted to exotic plant species<br />

rather than fostering native flora. Four <strong>of</strong> the Kiata colonies are<br />

within an 80ha Native Plant and Wildlife Reserve. Encroachment<br />

<strong>of</strong> exotic weeds on the site <strong>of</strong> the sole Castlemaine colony,<br />

which is not protected, may be a concern (Braby and Crosby in<br />

prep.). Sheep grazing has apparently reduced the number <strong>of</strong><br />

host plants in one degraded Kiata colony.<br />

<strong>Conservation</strong>: The need for conservation <strong>of</strong> P. p. lucida<br />

became both apparent and urgent because <strong>of</strong> the imminent<br />

subdivision <strong>of</strong> the major site on which the butterfly was<br />

fortuitously discovered. A briefing paper to the State Minister<br />

for <strong>Conservation</strong>, Forests and Land led to negotiations with the<br />

developers, who agreed to a moratorium on development until<br />

the feasibility <strong>of</strong> purchasing the site as a 'Butterfly Reserve'<br />

could be investigated. The total cost <strong>of</strong> this was projected at<br />

A$l million. The State Government contributed A$250,000<br />

and the Eltham Shire Council committed A$125,000. A highly<br />

organized public appeal during the next year raised a further<br />

$56,000 and during that time the 'Eltham Copper' became an<br />

important local emblem. The total <strong>of</strong> some A$426,000 is by far<br />

the largest sum ever committed to conservation <strong>of</strong> an invertebrate<br />

species/subspecies in Australia and a major part <strong>of</strong> the prime<br />

colony site (0.7ha) was purchased in early 1989. This was<br />

augmented substantially by the State Government transferring<br />

to the butterfly reserve an area (c. 2.6ha) <strong>of</strong> Education Department<br />

land adjacent to this which supported a further important<br />

colony.<br />

This case did much to increase public awareness <strong>of</strong> butterfly<br />

conservation and the general importance <strong>of</strong> invertebrates in the<br />

environment. Following initial assessment <strong>of</strong> conservation<br />

status (Crosby 1987), a management plan for P. p. lucida has<br />

been prepared (Vaughan 1987, 1988). The habitats reserved,<br />

though only a few hectares in extent, should be sufficient to<br />

sustain the populations at Eltham with adequate management.<br />

Management recommendations for these remnant urban<br />

populations include:<br />

i) protection <strong>of</strong> all existing colonies from threatening processes<br />

associated with urbanisation including: human activity<br />

(trampling, slashing or burning vegetation, dumping <strong>of</strong><br />

garbage, sullage overflow and changes to drainage patterns,<br />

potential overcollection); weed invasion; overgrowth <strong>of</strong><br />

food plants; and activities <strong>of</strong> other species <strong>of</strong> animals;<br />

167<br />

ii) provision for expansion <strong>of</strong> the habitat by prompting natural<br />

regeneration <strong>of</strong> Bursaria and propagation from seeds or<br />

cuttings, or transplanting plants from any sites to be destroyed<br />

by development;<br />

iii) provision for a ranger to foster practical management<br />

activities and monitor the effects <strong>of</strong> these.<br />

Listing <strong>of</strong> the subspecies under the Victorian Flora and<br />

Fauna Guarantee Act is controversial, although full provision<br />

for invertebrates is included in this pioneering legislation. A<br />

requirement <strong>of</strong> the Flora and Fauna Guarantee Act listing is the<br />

preparation <strong>of</strong> an Action Statement from the earlier management<br />

plan. This statement, detailing steps needed to ensure the<br />

butterfly's long-term survival, is at present in draft form (Webster<br />

1993). A further leaflet on the Eltham Copper has been issued<br />

recently to encourage its well being by habitat protection<br />

(Ahern 1993), and a Coordinating Group <strong>of</strong> scientists is<br />

overseeing the developing management plan. The Entomological<br />

Society <strong>of</strong> Victoria has placed P. p. lucida on its list <strong>of</strong><br />

butterflies to which a Voluntary Restricted Collecting Code<br />

applies.<br />

Current concerns include destruction <strong>of</strong> some <strong>of</strong> the small<br />

(unreserved) Eltham colonies and the coordination <strong>of</strong> effective<br />

measures to conserve the butterfly in other parts <strong>of</strong> Victoria.<br />

Attempts are being made, using the facilities <strong>of</strong> the 'Butterfly<br />

House' at the Royal Melbourne Zoological Gardens, to establish<br />

captive stock for future reintroduction to the wild.<br />

Acknowledgements<br />

M. Braby, D. Crosby and P. Vaughan are thanked for information<br />

additional to that included in the published reports.<br />

References<br />

AHERN, L. 1993. Eltham Copper Butterfly 'The People's choice'. Land for<br />

Wildlife Note No. 21. Department <strong>of</strong> <strong>Conservation</strong> and Natural Resources,<br />

Victoria.<br />

BRABY, M.F. 1990. The life history and biology <strong>of</strong> Paralucia pyrodiscus<br />

lucida Crosby (Lepidoptera: <strong>Lycaenidae</strong>).J. Aust. ent. Soc. 29: 41–51.<br />

BRABY, M.F. and CROSBY, D.F. In prep. The status and conservation <strong>of</strong> the<br />

Eltham Copper Butterfly, Paralucia pyrodiscus lucida Crosby, in Victoria,<br />

Australia.<br />

BRABY, M.F., CROSBY, D.F. and VAUGHAN, P.J. 1992. Distribution and<br />

range reduction in Victoria <strong>of</strong> the Eltham Copper butterfly Paralucia<br />

pyrodiscus lucida Crosby. Viet. Nat. 109: 154–161.<br />

CROSBY, D.F. 1951. A new geographical race <strong>of</strong> an Australian butterfly.<br />

Viet. Nat. 67: 225–227.<br />

CROSBY, D.F. 1987. The conservation status <strong>of</strong> the Eltham Copper Butterfly<br />

(Paralucia pyrodiscus lucida Crosby) (Lepidoptera: <strong>Lycaenidae</strong>). Arthur<br />

Rylah Institute for Environmental Research. Tech. Rpt. No. 8, Melbourne.<br />

VAUGHAN, P.J. 1987. (1988). The Eltham Copper Butterfly draft management<br />

plan. Arthur Rylah Institute for Environmental Research. Tech. Rpt. No.<br />

57, Melbourne.<br />

VAUGHAN, P.J. 1988. Management plan for the Eltham Copper Butterfly<br />

(Paralucia pyrodiscus lucida Crosby) (Lepidoptera: <strong>Lycaenidae</strong>). Arthur<br />

Rylah Institute for Environmental Research. Tech. Rpt. No. 79, Melbourne.<br />

WEBSTER, A. 1993. Eltham Copper Butterfly, Paralucia pyrodiscus lucida.<br />

Action Statement No. 39. (draft). Department <strong>of</strong> <strong>Conservation</strong> and Natural<br />

Resources, Victoria.


The Bathurst Copper, Paralucia spinifera Edwards and Common<br />

Country: Australia<br />

E.M. DEXTER and R.L. KITCHING<br />

Department <strong>of</strong> Ecosystem Management, University <strong>of</strong> New England, Armidale, NSW 2351, Australia<br />

Status and <strong>Conservation</strong> Interest: Status – endangered<br />

(Common and Waterhouse 1981).<br />

The species, endemic to Australia, is known only from three<br />

localities all within 50km <strong>of</strong> each other. The populations are<br />

highly fragmented and are subject to various threats including<br />

grazing, pasture improvements and establishment <strong>of</strong> forestry<br />

plantations.<br />

Taxonomy and Description: The Australian endemic genus<br />

Paralucia Waterhouse and Turner contains three species, P.<br />

spinifera, P. aurifera and P. pyrodiscus.<br />

The first specimen <strong>of</strong> P. spinifera was collected in October<br />

1964 at Yetholme (33°27'S 149°48'E) New South Wales, by<br />

I.F.B. Common and M.S. Upton. Despite subsequent searches,<br />

the species was not found again until October 1977 and it was<br />

<strong>of</strong>ficially described and named by Edwards and Common<br />

(1978). Apart from the difference in size, shape and colour <strong>of</strong><br />

the wing, Edwards and Common (1978) distinguished P.<br />

spinifera from its congeners P. aurifera (Blanchard) and P.<br />

pyrodiscus (Rosenstock) by its spine-like process that extends<br />

over the base <strong>of</strong> the tarsus on the tip <strong>of</strong> each fore tibia. Both<br />

sexes have this feature.<br />

Distribution: P. spinifera is known to occur at three main<br />

localities which are all within 50km <strong>of</strong> each other. The<br />

populations are highly fragmented both naturally, by occurring<br />

on ridges above 900m in altitude, and artificially, by pine<br />

plantations and areas <strong>of</strong> improved pasture.<br />

Population Size: Site A (see below) has the highest population<br />

<strong>of</strong> P. spinifera amounting to approximately 1120 individuals<br />

sighted over an eight-week period. Population sizes at the other<br />

two localities are smaller.<br />

Habitat and ecology: The three known localities for the<br />

species are remarkably uniform in geography, which suggests<br />

that the species has very specific microhabitat requirements.<br />

This has been confirmed by later studies (Dexter and Kitching<br />

unpublished).<br />

168<br />

All sites are on the edge <strong>of</strong> open eucalypt woodland (Specht<br />

1981) and usually on west to north-west facing slopes. The<br />

density <strong>of</strong> P. spinifera differs markedly in the three areas and is<br />

correlated with a spatial pattern <strong>of</strong> the larval food plant, native<br />

blackthorn Bursaria spinosa Cav. (Pittosporaceae).<br />

The butterfly has a mutualistic relationship with the ant<br />

Anonychomyrma itinerans. Ant surveys <strong>of</strong> known P. spinifera<br />

localities repeatedly yielded other ant species which may be<br />

important to the butterfly's life history. These were Iridomyrmex<br />

sp. 1,2,3, I. purpureus, I. rufoniger group 1 and 2, and<br />

Ochetellus sp.<br />

Ant surveys were also done in nearby areas, some only<br />

metres away, which appeared to be suitable habitat but lacked<br />

P. spinifera, and on every occasion the attendant ant was absent.<br />

One population which was high in numbers in 1989 crashed in<br />

1991. A follow-up survey showed that A. itinerans was not<br />

present.<br />

Athough A. itinerans has a wider distribution than P.<br />

spinifera, it too is restricted to regions above 900m in altitude.<br />

Apart from the central tablelands, A. itinerans is also found at<br />

Piccadilly Circus (35°22'S 148°49'E) (Australian Capital<br />

Territory), Barrington Tops (31°59'S 151°27'E) and Brown<br />

Mountain (36°36'S 149°23'E) in New South Wales.<br />

P. spinifera is univoltine and has a very early flight season<br />

which starts in the last week <strong>of</strong> September and finishes in mid-<br />

November. During this period the adult emergence is scattered<br />

and matings occur throughout the eight-week period.<br />

After mating, the female oviposits on bushes with ants,<br />

presumably detected using olfactory cues, and lays singletons<br />

or groups <strong>of</strong> up to four white eggs which darken to green with<br />

maturity. The egg group size <strong>of</strong> P. spinifera is considerably<br />

smaller than that <strong>of</strong> the closely related P. aurifera and P.<br />

pyrodiscus which are 15 and 12 eggs respectively (Braby<br />

1990). Eggs are laid on the lower third <strong>of</strong> the bush and are<br />

positioned on either the underside <strong>of</strong> leaves on the main trunk<br />

or on debris at the base <strong>of</strong> the plant. Eggs take approximately 15<br />

days to hatch. During the egg phase, the attendant ants are<br />

constantly searching the host plant, possibly seeking newly<br />

hatched larvae.<br />

After hatching, first instar larvae are immediately attended<br />

by one ant, which is curious as the early instar larvae do not have


ant-attracting organs. From instars one to three the larvae and<br />

ants are diurnal, feeding in both morning and afternoon and<br />

retreating to the nest at midday and dusk. After the third instar,<br />

the larvae and ants become nocturnal. The larvae <strong>of</strong> P. spinifera<br />

can have up to eight instars and can take between 60 and 70 days<br />

to pupate in laboratory conditions without ants.<br />

Preliminary laboratory studies (Dexter and Kitching,<br />

unpublished) have shown that larvae without ants remain on the<br />

bush permanently whilst larvae with ants return to the base <strong>of</strong><br />

the plant during daylight hours. Larvae reared without ants are<br />

also considerably smaller at pupation than larvae which have<br />

been reared with ants.<br />

Threats: The three main localities, which for the purposes <strong>of</strong><br />

this paper are termed sites A, B and C, differ greatly in the<br />

degree <strong>of</strong> habitat degradation and thus level <strong>of</strong> threat, which<br />

appears to be a reflection <strong>of</strong> the land tenure.<br />

Site A, which has the highest population <strong>of</strong> P. spinifera, is<br />

closest to the 'natural state' with a high level <strong>of</strong> plant diversity<br />

and a large community <strong>of</strong> ants present. This land is owned by<br />

the Department <strong>of</strong> Defence with restricted access, has not been<br />

burnt for 17 years, and has never been grazed by livestock. Site<br />

A also includes a nearby smaller site which is similar in<br />

condition to the main block described above.<br />

Site B supports five fragmented subpopulations and, although<br />

together the populations occupy an area <strong>of</strong> 5km 2 , a markrecapture<br />

study (Dexter and Kitching, unpublished) showed<br />

that the subpopulations do not intermingle and that the vagility<br />

<strong>of</strong> the adults is low. These sites are all on private land and have<br />

been or are still subject to grazing.<br />

Site C occurs within a nature reserve where it is fully<br />

protected by law. However, the population at this site is probably<br />

the most threatened <strong>of</strong> all. The site is heavily infested with<br />

blackberry, Rubus sp., is disturbed by feral pigs and threatened<br />

by fire control burns. Unfortunately the managing body is<br />

required to burn the area to reduce the risk <strong>of</strong> wildfire spreading<br />

from within the reserve to the neighbouring pine plantation.<br />

Unlike some lycaenids, for which grazing seems to be<br />

beneficial to the butterfly, grazing is detrimental to this species,<br />

as livestock trample and eat the seedlings <strong>of</strong> B. spinosa. This in<br />

turn inhibits juvenile recruitment <strong>of</strong> B. spinosa and can cause<br />

changes in the spatial pattern <strong>of</strong> the plant, leading to plant<br />

isolation. Larvae which occur on isolated bushes usually deplete<br />

their food supply so heavily that they starve or pupate<br />

prematurely. If larvae pupate at a smaller body size, the emerging<br />

adult will also be small and this can affect reproductive fitness<br />

because fecundity in butterflies is positively correlated with<br />

body size (Gilbert 1984).<br />

One sympathetic landholder is actively involved in managing<br />

her P. spinifera populations and provides hungry larvae with<br />

new bushes which ants and larvae colonise. It should be noted<br />

that for this management strategy to be successful the branches<br />

need to be entwined, as larvae do not seem to move across the<br />

ground.<br />

Overall the main threats to the survival <strong>of</strong> this species are:<br />

• ignorance <strong>of</strong> its existence and habitat degradation. The type<br />

169<br />

locality narrowly avoided being cleared during installation<br />

<strong>of</strong> a major power line, with no local appreciation <strong>of</strong> the<br />

butterfly's existence whatsoever.<br />

• grazing by cattle, sheep and goats, either by direct impact on<br />

the food plants or by use <strong>of</strong> pasture fertilisation to improve<br />

fodder quality.<br />

• clearing for establishment <strong>of</strong> forestry plantations. Along<br />

with pasture improvement, this process has probably<br />

produced the current highly fragmented distribution <strong>of</strong> the<br />

species in the area, although it must always have been a very<br />

local species. Clearing and habitat change were <strong>of</strong> course<br />

carried out in ignorance <strong>of</strong> the existence <strong>of</strong> the species.<br />

• all sites are under some threat from exotic weeds, the most<br />

significant being blackberry (Rubus spp.) and scotch broom<br />

(Cytisus scoparius). At one site, the Bursaria is in danger <strong>of</strong><br />

being completely overgrown by the broom and at the only<br />

protected site (Winburndale Nature Reserve) uncontrolled<br />

spread <strong>of</strong> blackberries is a threat to part <strong>of</strong> the site at least.<br />

<strong>Conservation</strong>: Paralucia spinifera encapsulates many <strong>of</strong> the<br />

problems associated with the conservation <strong>of</strong> lycaenids. The<br />

scatter <strong>of</strong> highly restricted sites each with a different set <strong>of</strong><br />

actual and potential problems and the relatively small changes<br />

that would be required to actually eliminate whole populations<br />

make generalisations difficult. Crucial to any management <strong>of</strong><br />

the species is local awareness and sympathy and that has been<br />

created over the last few years. Only those who have regular,<br />

albeit casual, contact with the sites concerned can monitor day<br />

to day changes.<br />

Some general threats, particularly overgrowth by noxious<br />

weeds and damage to hostplants by feral animals, should be<br />

controlled by application <strong>of</strong> existing programmmes in National<br />

Parks and other public lands. Sites on private land need to be<br />

monitored for these impacts and the impact <strong>of</strong> stock, particularly<br />

goats, which may browse upon the host plants, and cattle which<br />

can produce soil impaction and other disturbance inimical to ant<br />

populations.<br />

Major land use changes involving clearing for conifer<br />

plantations and fertilisation for pasture 'improvement' are<br />

probably responsible for the present limited distribution <strong>of</strong> the<br />

species and obviously any further changes present additional<br />

threats. The occurrence <strong>of</strong> some butterfly populations on<br />

protected land is encouraging although other sites need to be<br />

monitored in this connection. One site is being proposed for<br />

listing under National Estate regulations because <strong>of</strong> the presence<br />

<strong>of</strong> this butterfly.<br />

There is circumstantial evidence for one site that<br />

overcollecting has severely reduced the population. The short<br />

flight period, very restricted areas in which populations occur,<br />

and a tendency by collectors to take long series for no justifiable<br />

reason may have contributed to making this species more<br />

sensitive than most to collecting pressures. Lepidopterists have<br />

argued vigorously for a self-regulatory approach to butterfly<br />

conservation rather than one driven by species-specific<br />

legislation. Collectors need to be very aware <strong>of</strong> the sensitivity<br />

<strong>of</strong> species like Paralucia spinifera for which self-regulation


does not appear to have worked. The case could certainly<br />

be made for total protection <strong>of</strong> the species.<br />

References<br />

BRABY, M.F. 1990. The life history and biology <strong>of</strong> Paralucia<br />

pyrodiscuslucida Crosby (Lepidoptera: <strong>Lycaenidae</strong>).J. Aust. ent.<br />

Soc. 29:41–51.<br />

170<br />

COMMON, I.F.B. and WATERHOUSE, D.F. 1981. <strong>Butterflies</strong> <strong>of</strong> Australia.<br />

Angus and Robertson, Sydney.<br />

EDWARDS, E.D. and COMMON, I.F.B. 1978. A new species <strong>of</strong> Paralucia<br />

Waterhouse and Turner from New South Wales (Lepidoptera: <strong>Lycaenidae</strong>).<br />

Aust. ent. Mag. 5 (4): 65–70.<br />

GILBERT, N. 1984. Control <strong>of</strong> fecundity in Pieris rapae I. The problem. J.<br />

Anim. Ecol. 53: 581–588.<br />

SPECHT, R.L. 1981. Foliage projective cover and standing biomass. In:<br />

Gillison, A.N. and Anderson, D.J. (Eds). Vegetation classification in<br />

Australia. CSIRO and ANU Press, Canberra.


The Australian Hairstreak, Pseudalmenus chlorinda (Blanchard)<br />

G.B. PRINCE<br />

Department <strong>of</strong> Parks, Wildlife and Heritage, 134 Mrs. Macquarie's Road, Hobart, Tasmania 7001, Australia<br />

Country: Australia.<br />

Status and <strong>Conservation</strong> Interest: Status – Mainland<br />

subspecies: P. c. chloris, rare; P. c. fisheri and P.c.<br />

barringtonensis, vulnerable; P. c. zephyrus, not threatened.<br />

Tasmanian subspecies: P. c. conara and P. c. chlorinda,<br />

vulnerable (both indeterminate: Red List); P. c. myrsilus and<br />

the un-named P. c. zephyrus-likc form, endangered.<br />

Several <strong>of</strong> the subspecies <strong>of</strong> P. chlorinda are extremely<br />

localised, particularly in Tasmania where most conservation<br />

attention to this species has been paid. Several forms <strong>of</strong> the<br />

Hairstreak in Tasmania were stated by Couchman and Couchman<br />

(1977) 'to have suffered more than any local species <strong>of</strong> butterfly',<br />

and much suitable habitat has been destroyed. The Couchmans<br />

expressed grave concern for the butterfly's future in Tasmania,<br />

and a more recent survey (Prince 1988) also stressed the need<br />

for conservation measures. Many colonies have become extinct,<br />

and most <strong>of</strong> the remainder face threats to their continued well<br />

being. The species as a whole in Tasmania is considered<br />

vulnerable.<br />

Taxonomy and Description: P. chlorinda is the sole species <strong>of</strong><br />

the endemic Australian thecline genus Pseudalmenus Druce,<br />

and is known from southeastern mainland Australia and<br />

Tasmania. It shows substantial geographical variation (Figure<br />

1) and seven named subspecies are widely recognised (Common<br />

and Waterhouse 1981), together with at least one other<br />

Tasmanian form. Couchman and Couchman (1977) noted<br />

another two Tasmanian forms which were by then extinct.<br />

Collectively, these forms constitute one <strong>of</strong> the most diverse<br />

polytypic <strong>Lycaenidae</strong> in Australia, probably representing<br />

incipient speciation, and <strong>of</strong> considerable evolutionary interest.<br />

Distribution: All eight subspecies have highly circumscribed<br />

distributions (see Figure 1) and only one, P. c. zephyrus, can be<br />

considered to be reasonably widespread. P. c. chloris is regarded<br />

as rare and local in New South Wales and the other mainland<br />

subspecies, P. c. fisheri and P. c. barringtonensis , are highly<br />

localised in the Grampians Mountains, Victoria and Barrington<br />

Tops, New South Wales, respectively. Of the Tasmanian taxa,<br />

P. c. chlorinda is found from Hobart to north <strong>of</strong> Swansea and<br />

171<br />

westward to the South Esk and Upper Tamar Valleys, P. c.<br />

conara is known from the midlands and P. c. myrsilus from the<br />

Tasman and Forestier Peninsulas and a small part <strong>of</strong> the east<br />

coast. The un-namedP. c. zephyrus-like form is known from the<br />

northeast <strong>of</strong> Tasmania. The recent discovery <strong>of</strong> the species in<br />

western Tasmania suggests the likelihood that other colonies<br />

may exist in that remote area: that race has been referred<br />

tentatively to P. c. chlorinda, but specimens are not available.<br />

Population Size: Most subspecies are clearly very localised<br />

but little information is available on population size, especially<br />

for the mainland taxa. Many <strong>of</strong> the Tasmanian sites surveyed by<br />

Prince (1988) had very low 'occupation rates' <strong>of</strong> suitable<br />

eucalypt trees, with few hairstreak pupae recorded from them.<br />

Many populations may, indeed, be confined to single eucalypt<br />

trees in very small isolated patches <strong>of</strong> habitat, from which<br />

recolonisation is probably unlikely. Detailed information on<br />

Pseudalmenus dispersal is not available, but it is believed to be<br />

largely sedentary in habit.<br />

Habitat and Ecology: Eggs are laid singly or in small groups<br />

on young twigs <strong>of</strong> a larval foodplant, Acacia. In Tasmania, the<br />

usual foodplant is the bipinnate A. dealbata (80% <strong>of</strong> Prince's<br />

records), with the closely related A. mearnsii (9%) also utilised.<br />

Other acacias, particularly the phyllodinous A. melanoxylon,<br />

have also been recorded, more especially for the mainland<br />

subspecies. Larvae feed on the Acacia foliage and are attended<br />

by small, black ants (Iridomyrmex foetans). The role <strong>of</strong> the ants<br />

is not clear, but they appear to be essential, and probably protect<br />

the caterpillars from parasitoids. Caterpillars pupate under the<br />

loose bark <strong>of</strong> nearby eucalypts, predominantly Eucalyptus<br />

viminalis (92% <strong>of</strong> pupae) in Tasmania, growing within a few<br />

metres <strong>of</strong> the acacias. Most (77%) grew within 2m, and few<br />

pupae were found on eucalypts only 10m from larval foodplants.<br />

The most frequently selected eucalypts were 15–18m tall, with<br />

diameters <strong>of</strong> 75–150cm. Pupae occurred up to around 2m from<br />

the ground, and most were on the northern half <strong>of</strong> the tree.<br />

Adults fly in spring and early summer, with slight differences<br />

in flight period between subspecies, and most forms are<br />

univoltine. A partial second generation has been suggested for<br />

P. c. fisheri (Common and Waterhouse 1981) but otherwise the


Figure 1. Distribution <strong>of</strong> subspecies <strong>of</strong> Pseudalmenus chlorinda in<br />

southeastern Australia, with details <strong>of</strong> Tasmanian subspecies. Mainland<br />

data after Common and Waterhouse (1981).<br />

period from December to at least August or September is passed<br />

as pupae. This stage therefore constitutes a suitable one for<br />

monitoring population sizes <strong>of</strong> Pseudalmenus. Pupae are <strong>of</strong>ten<br />

on trees occupied by I. foetans, and the scent <strong>of</strong> the ants might<br />

influence selection <strong>of</strong> host trees (Prince 1988). The ant can also<br />

be abundant on trees lacking Pseudalmenus.<br />

Threats: At many sites in Tasmania where P. chlorinda has<br />

become extinct, the habitat has been destroyed or severely<br />

disturbed. Pastoral clearing (52%) and fire (direct implication<br />

in loss <strong>of</strong> 24% <strong>of</strong> sites, contributory to the loss <strong>of</strong> a further 20%)<br />

were the predominant factors involved, and grazing, forestry<br />

(clearing for timber and woodchips) and housing subdivisions<br />

172<br />

accounted for the remainder. For extant colonies, Prince (1988)<br />

assessed fire as the most common threat (31 %) though clearing<br />

and subdivision are also significant.<br />

Pupae were recorded in the mid-1980s at only 15 <strong>of</strong> the 43<br />

sites noted by earlier workers, but some new sites were also<br />

found. Of 66 locations recorded for P. chlorinda in Tasmania,<br />

only 12 <strong>of</strong> the 40 which still support the hairstreak were<br />

considered to be reasonably secure (Prince 1988). P. c. myrsilus<br />

is especially restricted, and P. c. near zephyrus was found in<br />

only four areas. Most known sites are small and threatened. At<br />

many, pupae were found only on a single tree. Couchman and<br />

Couchman (in Prince 1988) observed instances where a single<br />

tree had supported a population <strong>of</strong> P. chlorinda for many years,<br />

and that population had disappeared once the tree had been cut<br />

down.<br />

<strong>Conservation</strong>: Couchman and Couchman (1977) emphasised<br />

the threefold needs <strong>of</strong> P. chlorinda: the close associations <strong>of</strong> a<br />

suitable Eucalyptus; a suitable Acacia; and the specific ant.<br />

They emphasised that the destruction <strong>of</strong> any <strong>of</strong> these might lead<br />

to the butterfly's local extermination. Protection <strong>of</strong> these<br />

resources in suitable habitats is clearly the major conservation<br />

need for P. chlorinda, both in terms <strong>of</strong> reservation <strong>of</strong> habitat and<br />

management within reserves and on private land. Many <strong>of</strong> the<br />

populations are in reserved sites, but some <strong>of</strong> these have been<br />

subject to clearing and fire. Protection <strong>of</strong> habitats on private<br />

land may be feasible in some instances, but sites close to urban<br />

centres are under considerable pressure from urban expansion.<br />

Increased community awareness <strong>of</strong> the butterfly is also needed.<br />

Active management, for example guarding against clearing<br />

<strong>of</strong> acacias near eucalypts, merits promotion, and a trial reported<br />

by Prince (1988) suggests that translocation <strong>of</strong> the butterfly to<br />

new suitable habitats might be feasible as a management tool.<br />

Studies <strong>of</strong> the status <strong>of</strong> mainland subspecies are also a priority<br />

for this species, together with clarification <strong>of</strong> the status <strong>of</strong><br />

populations in western Tasmania.<br />

References<br />

COMMON, I.F.B. and WATERHOUSE, D.F. 1981. <strong>Butterflies</strong> <strong>of</strong> Australia.<br />

(2nd edn.) Angus & Robertson, Sydney.<br />

COUCHMAN, L.E. and COUCHMAN, R. 1977. The butterflies <strong>of</strong> Tasmania.<br />

Tasmanian Year Book (1977): 66–96.<br />

PRINCE, G.B. 1988. The conservation status <strong>of</strong> the Hairstreak Butterfly<br />

Pseudalmenus chlorinda Blanchard in Tasmania (report to Department <strong>of</strong><br />

Lands, Parks and Wildlife, Tasmania).


APPENDIX 1<br />

<strong>IUCN</strong> Red Data Book categories<br />

Many <strong>of</strong> the species discussed in this book have been allocated,<br />

sometimes tentatively, to the traditional <strong>IUCN</strong> Red Data Book<br />

categories. They are subject to revision <strong>of</strong> status as new<br />

information is incorporated. The categories are defined as<br />

follows:<br />

Extinct (Ex)<br />

Species not definitely located in the wild during the past 50<br />

years (criterion as used in the Convention on International<br />

Trade in Endangered Species <strong>of</strong> Wild Fauna and Flora). (For<br />

some butterflies, the status is applied to taxa which are known<br />

to have become extinct by, for example, destruction <strong>of</strong> the last<br />

or only known colony, without regard to the 50 year period:<br />

TRN.)<br />

Endangered (E)<br />

Taxa in danger <strong>of</strong> extinction and whose survival is unlikely if<br />

the causal factors continue to operate. These include species<br />

whose habitats or numbers have been reduced drastically and<br />

may be in danger <strong>of</strong> imminent extinction. The category also<br />

includes taxa which are probably already extinct, but which<br />

have been seen during the last 50 years (see comment in<br />

parentheses, above: 'Endangered' implies some uncertainty<br />

over whether or not the species is extinct: TRN).<br />

Vulnerable (V)<br />

Taxa believed likely to become Endangered in the near future<br />

if causal factors continue to operate. A wide range <strong>of</strong> threats is<br />

173<br />

included, from habitat destruction to overexploitation, and<br />

other environmental disturbances. The taxa need not, necessarily,<br />

be rare if threats operate over a large range.<br />

Rare (R)<br />

Taxa with small world populations which are not at present<br />

Endangered or Vulnerable, but are at risk. They are usually in<br />

limited geographical areas or habitats, or in low numbers over<br />

a more extensive range.<br />

Indeterminate (I)<br />

Taxa known to be Endangered, Vulnerable, or Rare, but for<br />

which there is insufficient information to determine which <strong>of</strong><br />

these is the appropriate category.<br />

Out <strong>of</strong> Danger (O)<br />

Taxa formerly in one <strong>of</strong> the above categories but which are now<br />

considered relatively secure because <strong>of</strong> effective conservation<br />

measures and/or removal <strong>of</strong> previous threats to their survival.<br />

'Threatened' is a general term to denote species which are<br />

Endangered, Vulnerable, Rare, or Indeterminate.<br />

A 'Threatened Community' is a group <strong>of</strong> ecologically linked<br />

taxa occurring within a defined area, which are all under the<br />

same threat and require similar conservation measures.


Other Occasional Papers <strong>of</strong> the <strong>IUCN</strong> Species Survival Commission<br />

1. Species <strong>Conservation</strong> Priorities in the Tropical Forests <strong>of</strong> Southeast Asia. Edited by R.A. Mittermeier and W.R.<br />

Constant, 1985, 58pp, £7.50, US$15.00. (Out <strong>of</strong> print)<br />

2. Priorités en matière de conservation des espèces à Madagascar. Edited by R.A. Mittermeier, L.H. Rakotovao, V.<br />

Randrianasolo, E.J. Sterling and D. Devitre, 1987, 167pp, £7.50, US$15.00.<br />

3. <strong>Biology</strong> and <strong>Conservation</strong> <strong>of</strong> River Dolphins. Edited by W.F. Perrin, R.K. Brownell, Zhou Kaiya and Liu Jiankang,<br />

1989, 173pp, £10.00, US$20.00.<br />

4. Rodents. A World Survey <strong>of</strong> Species <strong>of</strong> <strong>Conservation</strong> Concern. Edited by W.Z. Lidicker, Jr., 1989, 60pp, £7.50,<br />

US$15.00.<br />

5. The <strong>Conservation</strong> <strong>Biology</strong> <strong>of</strong> Tortoises. Edited by I.R. Swingland and M.W. Klemens, 1989, 202pp, £12.50,<br />

US$25,00.<br />

6. Biodiversity in Sub-Saharan Africa and its Islands, 1991, £12.50, US$25.00.<br />

7. Polar Bears: Proceedings <strong>of</strong> the Tenth Working Meeting <strong>of</strong> the <strong>IUCN</strong>/SSC Polar Bear Specialist Group, 1991,<br />

£10.00, US$20.00.<br />

Where to order:<br />

<strong>IUCN</strong> Publications Services Unit, 219 Huntingdon Road, Cambridge, CB3 0DL, U.K. Please pay by cheque/<br />

international money order to <strong>IUCN</strong>. Add 15% for packing and surface mail costs. A catalogue <strong>of</strong> <strong>IUCN</strong> publications can<br />

be obtained from the above address.


<strong>IUCN</strong> Species Survival Commission<br />

The Species Survival Commission (SSC) is one <strong>of</strong> six volunteer commissions <strong>of</strong> <strong>IUCN</strong> – The World <strong>Conservation</strong> Union,<br />

a union <strong>of</strong> sovereign states, government agencies and non-governmental organizations. <strong>IUCN</strong> has three basic<br />

conservation objectives: to secure the conservation <strong>of</strong> nature, and especially <strong>of</strong> biological diversity, as an essential<br />

foundation for the future; to ensure that where the earth's natural resources are used this is done in a wise, equitable and<br />

sustainable way; and to guide the development <strong>of</strong> human communities towards ways <strong>of</strong> life that are both <strong>of</strong> good quality<br />

and in enduring harmony with other components <strong>of</strong> the biosphere.<br />

The SSC's mission is to conserve biological diversity by developing and executing programs to save, restore and wisely<br />

manage species and their habitats. A volunteer network comprised <strong>of</strong> 4,800 scientists, field researchers, government<br />

<strong>of</strong>ficials and conservation leaders from 169 countries, the SSC membership is an unmatched source <strong>of</strong> information about<br />

biological diversity and its conservation. As such, SSC members provide technical and scientific counsel for conservation<br />

projects throughout the world and serve as resources to governments, international conventions and conservation<br />

organizations.<br />

The <strong>IUCN</strong>/SSC Occasional Paper series focuses on a variety <strong>of</strong> conservation topics including conservation overviews on<br />

a regional or taxonomic basis and proceedings <strong>of</strong> important meetings.<br />

<strong>IUCN</strong>/SSC also publishes an Action Plan series that assesses the conservation status <strong>of</strong> species and their habitats, and<br />

specifies conservation priorities. The series is one <strong>of</strong> the world's most authoritative sources <strong>of</strong> species conservation<br />

information available to nature resource managers, conservationists and government <strong>of</strong>ficials around the world.<br />

Published by <strong>IUCN</strong><br />

This book is part <strong>of</strong> The <strong>IUCN</strong> <strong>Conservation</strong> Library<br />

For a free copy <strong>of</strong> the complete catalogue please write to<br />

<strong>IUCN</strong> Publications Services Unit,<br />

219 Huntingdon Road, Cambridge CB3 0DL, U.K.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!