20.01.2013 Views

Comparative Parasitology 67(2) 2000 - Peru State College

Comparative Parasitology 67(2) 2000 - Peru State College

Comparative Parasitology 67(2) 2000 - Peru State College

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

July <strong>2000</strong> Number 2<br />

<strong>Comparative</strong> <strong>Parasitology</strong><br />

Formerly the<br />

Journal of the Helminthological Society of Washington<br />

CONTENTS<br />

KRITSKY, D. C., F. A. JIMENEZ-RUIZ, AND O. SEY. Diplectanids (Monogenoidea: Dactylogyridea)<br />

from the Gills of Marine Fishes of the Persian Gulf off Kuwait 145<br />

DAILEY, M. D., AND S. R. GOLDBERG. Langeronia burseyi sp. n. (Trematoda: Lecithodendriidae)<br />

from the California Treefrog, Hyla cadaverina (Anura: Hylidae), with<br />

Revision of the Genus Langeronia Caballero and Bravo-Hollis, 1949 165<br />

BOUAMER, S., AND S. MoRAND. Oxyuroids of Palearctic Testudinidae: New Definition<br />

'of the Genus Thaparia Ortlepp, 1933 (Nematoda: Pharyngodonidae), Redescription<br />

of Thaparia thapari thapari, and Descriptions of Two New Species 169<br />

MUZZALL, P. M. Parasites of Farm-Raised Trout in Michigan, U.S.A. . 181<br />

BULLARD, S. A., G. W. BENZ, R. M. OVERSTREET, E. H. WILLIAMS, JR., AND J. HEMDAL.<br />

Six New Host Records and an Updated List of Wild Hosts for Neobenedenia melleni<br />

(MacCallum) (Monogenea: Capsalidae) . _<br />

DUCLOS, L. M., AND D. J. RICHARDSON. Hymenolepis nana in Pet Store Rodents<br />

BOLEK, M. G., AND J. R. COGGINS. Seasonal Occurrence and Community Structure of<br />

Helminth Parasites from the Eastern American Toad, Bufo americanus americanus,<br />

from Southeastern Wisconsin, U.S.A 202<br />

MACHADO, P. M., S. C. DE ALMEIDA, G. C. PAVANELLI, AND R. M. TAKEMOTO. Ecological<br />

Aspects of Endohelminths Parasitizing Cichla monoculus Spix, 1831 (Perciformes:<br />

Cichlidae) in the Parana River near Porto Rico, <strong>State</strong> of Parana, Brazil 210<br />

LYONS, E. T, T. R. SPRAKER, K. D. OLSON, S. C. TOLLIVER, AND H. D. BAIR. Prevalence<br />

of Hookworms (Uncinaria lucasi Stiles) in Northern Fur Seal (Callorhinus ursinus<br />

Linnaeus) Pups on St. Paul Island, Alaska, U.S.A.: 1986-1999 218<br />

HASEGAWA, H., T. Doi, A. FUJISAKI, AND A. MIYATA. Life History of Spiroxys hanzaki<br />

Hasegawa, Miyata, et Doi, 1998 (Nematoda: Gnathostomatidae) 224<br />

BOCZON, K., AND B. WARGIN. Inducible Nitric Oxide Synthase in the Muscles of Trichinella<br />

sp.-Infected Mice Treated with Glucocorticoid Methylprednisolone 230<br />

FUJINO, T., T. SHINOHARA, K. FUKUDA, H. ICHIKAWA, T. NAKANO, AND B. FRIED. The<br />

Expulsion of Echinostoma trivolvis: Worm Kinetics and Intestinal Cytopathology<br />

in Jirds, Meriones unguiculatus „ •. 236<br />

DARAS, M. R., S. SISBARRO, AND B. FRIED. Effects of a High-Carbohydrate Diet on<br />

Growth of Echinostoma caproni in ICR Mice _ . 241<br />

KOGA, M., H. AKAHANE, R. LAMOTHE-ARGUMEDO, D. OSORIO-SARABIA, L. GARC{A-PRIETO,<br />

J. M. MARTINEZ-CRUZ, S. P. DlAZ-CAMACHO, AND K. NODA. Surface Ultrastructure<br />

of Larval Gnathostoma cf. binucleatum from Mexico ... ..... 244<br />

(Continued on Outside Back Cover)<br />

Copyright © 2011, The Helminthological Society of Washington


THE HELMINTHOLOGICAL SOCIETY OF WASHINGTON<br />

THE SOCIETY meets approximately five times per year for the presentation and discussion of papers<br />

in any and all branches of parasitology or related sciences. All interested persons are invited to attend.<br />

Persons interested in membership in the Helminthological Society of Washington may obtain application<br />

blanks in recent issues of COMPARATIVE PARASITOLOGY. A year's subscription to<br />

COMPARATIVE PARASITOLOGY is included in the annual dues of $25.00 for domestic membership<br />

and $28.00 for foreign membership. Institutional subscriptions are $50.00 per year. Applications for<br />

membership, accompanied by payments, may be sent to the Corresponding Secretary-Treasurer, Nancy<br />

D. Pacheco, 9708 DePaul Drive, Bethesda, MD 208(17, U:S,A.<br />

The HelmSoc internet home page is located at http://www.gettysburg.edu/~shendrix/helmsoc.html<br />

OFFICERS OF THE SOCIETY FOR <strong>2000</strong><br />

President: DENNIS J. RICHARDSON<br />

Vice President: LYNN K. CARTA<br />

Corresponding Secretary-Treasurer: NANCY D. PACHECO<br />

Recording Secretary: W. PATRICK CARNEY<br />

Archivist/Librarian: PATRICIA A. PILITT<br />

Custodian of Back Issues: J. RALPH LICHTENFELS<br />

Representative to the American Society of Parasitologists: ERIC P. HOBERG<br />

Executive Committee Members-at-Large: RALPH P. ECKERLIN, <strong>2000</strong><br />

WILLIAM E. MOSER, <strong>2000</strong><br />

ALLEN L.RICHARDS, 2001<br />

BENJAMIN M. ROSENTHAL, 2001<br />

Immediate Past President: ERIC P. HOBERG<br />

COMPARATIVE PARASITOLOGY<br />

COMPARATIVE PARASITOLOGY is published semiannually at Lawrence, Kansas by the<br />

Helminthological Society of Washington. Papers need not be presented at a meeting to be published in<br />

the journal. Publication of COMPARATIVE PARASITOLOGY is supported in part by the Brayton H.<br />

Ransom Memorial Trust Fund.<br />

MANUSCRIPTS should be sent to the EDITORS, Drs. Willis A. Reid, Jr., and Janet W. Reid, 6210<br />

Hollins Drive, Bethesda, MD 20817-2349, email: jwrassoc@erols.com. Manuscripts must be typewritten,<br />

double spaced, and in finished form. Consult recent issues of the journal for format and style. The<br />

original and three copies are required. Photocopies of drawings may be submitted for review purposes<br />

but glossy prints of halftones are required; originals will be requested after acceptance of the manuscript.<br />

Papers are accepted with the understanding that they will be published only in the journal.<br />

REPRINTS may be ordered from the PRINTER at the same time the corrected proof is returned to<br />

the EDITORS.<br />

AUTHORS' CONTRIBUTIONS to publication costs (currently S50/page for members, $100/page<br />

for non-members) will be billed by Allen Press and are payable to the Society.<br />

BACK ISSUES and VOLUMES of COMPARATIVE PARASITOLOGY are available, Inquiries concerning<br />

back volumes and current subscriptions should be directed to the business office.<br />

BUSINESS OFFICE. The Society's business office is at Lawrence, Kansas. All inquiries concerning<br />

subscriptions or back issues and all payments for dues, subscriptions, and back issues should be<br />

addressed to: Helminthological Society of Washington, % Allen Press, Inc., 1041 New Hampshire St.,<br />

Lawrence, Kansas 66044, U.S.A.<br />

<strong>2000</strong><br />

ROY C.-ANDERSON<br />

RALPH P. ECKERLIN<br />

ROBIN N. HUETTEL<br />

FUAD M. NAHHAS<br />

DANNY B. PENCE<br />

JOSEPH F. URBAN<br />

EDITORIAL BOARD<br />

WILLIS A. REID, JR. & JANET W. REID, Editors<br />

2001<br />

WALTER A BOEGER<br />

WILLIAM F. FONT<br />

DONALD FORRESTER<br />

J. RALPH LICHTENFELS<br />

JOHN S. MACKIEWICZ<br />

BRENT NICKOL<br />

© The Helminthological Society of Washington <strong>2000</strong><br />

ISSN 1049-233X<br />

2002<br />

DANIEL R. BROOKS<br />

HIDEO HASEGAWA<br />

SHERMAN S. HENDRIX<br />

JAMES E. JOY<br />

DAVID MARCOGLIESE<br />

DANTE S. ZARLENGA<br />

This paper meets the requirements of ANSI/NISO Z39.48-1992 (Permanence of Paper).<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 145-164<br />

Diplectanids (Monogenoidea: Dactylogyridea) from the Gills of<br />

Marine Fishes of the Persian Gulf off Kuwait<br />

DELANE C. KRITSKY,M F. AGUSTIN JiMENEZ-Ruiz,2 AND OTTO SEY3<br />

1 Department of Health and Nutrition Sciences, <strong>College</strong> of Health Professions, Box 8090, Idaho <strong>State</strong><br />

University, Pocatello, Idaho 83209, U.S.A. (e-mail: kritdela@isu.edu),<br />

2 Laboratorio de Helmintologfa, Institute de Biologia, National Autonomous University of Mexico (UNAM),<br />

Apartado Postal 70-153, Ciudad de Mexico D.F., Mexico, and<br />

3 Department of Zoology, University of Kuwait, P.O. Box 5969, Safat 13060, Kuwait; current address: H-7633<br />

Pecs, Ercbanyasz u. 10, Hungary<br />

ABSTRACT: Seventeen species of Diplectanidae were collected from the gills of 17 species of marine fishes from<br />

the Persian Gulf off Kuwait. Lepidotrema kuwaitcnsis sp. n. from Terapon puta (Teraponidae), Lamellodiscus<br />

furcillatus sp. n. from Diplodus noct (Sparidae), and Protolamellodiscus senilobatus sp. n. from Argyrops spinifer<br />

and A. filamentosus (Sparidae) are described. Diplectanum cazauxi from Sphyraena jello and S. obtusata (Sphyraenidae)<br />

(new host and geographic records), D. sillagonum from Sillago siharna (Sillaginidae) (new geographic<br />

record), Pseudolarnellodiscus sphyraenae from Sphyraena chrysotaenia (Sphyraenidae) (new host and geographic<br />

records), and Calydiscoides flexuosus from Nemipterus peronii and N. bipunctatus (Nemipteridae) (new host<br />

and geographic records) are redescribed. An incidental geographic record for C. flexuosus on N. japonicus from<br />

the western coast of India is included. Ten diplectanid species from 8 hosts were unidentified for lack of sufficient<br />

specimens. Diplectanum longipenis (synonym: Squamodiscus longipenis) is transferred to Lepidotrema. Squamodiscus<br />

is removed from synonymy with Diplectanum and becomes a junior subjective synonym of Lepidotrema.<br />

Calydiscoides indianus (synonyms: Lamellospina Indiana and C. indicus) is a junior subjective synonym<br />

of C. flexuosus.<br />

KEY WORDS: Monogenoidea, monogenean, Diplectanidae, Calydiscoides flexuosus, Diplectanum cazauxi, Diplectanum<br />

sillagonum, Diplectanum sp., Lamellodiscus furcillatus sp. n., Lamellodiscus sp., Lepidotrema kuwaitensis<br />

sp. n., Lepidotrema longipenis comb, n., Protolamellodiscus senilobatus sp. n., Pseudolamellodiscus<br />

sphyraenae, Pseiidorhabdosynochus sp., Acanthopagrus berda, Acanthopagrus bifasciatus, Acanthopagrus latus,<br />

Argyrops filamentosus, Argyrops spinifer, Diplodus noct, Epinephelus areolatus, Epinephelus tauvina, Hemiramphus<br />

marginatus, Nemipterus bipunctatus, Nemipterus peronii, Otolithes argenteus, Sillago sihama, Sphyraena<br />

chrysotaenia, Sphyraena jello, Sphyraena obtusata, Terapon puta, Persian Gulf, Kuwait.<br />

A survey of the helminth parasites infesting described by Kritsky et al. (1986). Measurements, all<br />

marine fishes off the Kuwaiti coast by O.S. was in njicrometers, were made with a filar micrometer ac-<br />

_. . *nn~ ?n i- cording to procedures of Mizelle and Klucka (1953);<br />

conducted between October 1992 and December average measurements are followed by ranges and<br />

1996. Species of Diplectanidae (Monogenoidea) number (n) of specimens measured in parentheses; unwere<br />

found on the gills of 17 marine fishes rep- stained flattened specimens mounted in Gray and<br />

resenting the Hemiramphidae, Nemipteridae, Wess' medium were used to obtain measurements of<br />

Sciaenidae, Serranidae, Sillaginidae, Sparidae, the hap'oral sclerites and copulatory complex; other<br />

„ . . , , m • j rr,, .<br />

Sphyraemdae, and Teraponidae. The present pa-<br />

^ J f . .<br />

per includes descriptions and taxonomic considmeasurements<br />

were obtained from unflattened specit<br />

. , . „ ., t . , . v, .<br />

mens stained in Gomon s tnchrome and mounted in<br />

Canada balsam; the dimension of the pyriform ovary<br />

erations of 3 new and 5 previously described is the greatest width. Numbering of hook pairs follows<br />

species. the scheme proposed by Mizelle (1936; see Mizelle<br />

and Price, 1963). Type specimens of new species and<br />

Materials and Methods voucher specimens of previously described species<br />

, . , . i i , /- . i . «•<br />

Hosts were obtained from the local fish market, Ku-<br />

, , ,. . ,. , . . . . ' .<br />

wait, and examined directly tor helminth parasites. Diplectanids<br />

were removed from the gills of respective<br />

were deposited in the United <strong>State</strong>s National Parasite<br />

_, „ . /T TOXT~,x r» i* -n A* , ^ ^ ,<br />

Collection (USNPC), Beltsville, Maryland, and the<br />

, , - , ,, • r- i. TT • • AT i , n<br />

helminth collection of the University of Nebraska <strong>State</strong><br />

hosts, fixed, and stored as described by Sey and Nah- Museum (HWML), Lincoln, Nebraska, U.S.A., as mhas<br />

(1997); vials containing the helminths were then dlcated m the respective species accounts. For cornshipped<br />

to Idaho <strong>State</strong> University. Methods of staining, Paratlve Purposes, the following specimens were exmounting,<br />

and illustration of diplectanids were those ammed: 3 voucher specimens of Lepidotrema tenue<br />

Johnston and Tiegs, 1922 (USNPC 63156); 4 voucher<br />

specimens of Lepidotrema bidyana Murray, 1931<br />

4 Corresponding author. (USNPC 63157); 5 voucher specimens of Lepidotrema<br />

145<br />

Copyright © 2011, The Helminthological Society of Washington


146 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

angusta (Johnston and Tiegs, 1922) (USNPC 63158);<br />

holotype, 22 paratypes of Pseudolamellodiscus sphyraenae<br />

Yamaguti, 1953 (Meguro Parasitological Museum,<br />

Tokyo, Japan [MPM] 22556); holotype, numerous<br />

paratypes of Lamellodiscus convolutus Yamaguti,<br />

1953 (MPM 22558); holotype, 27 paratypes of Lamellodiscus<br />

flexuosus Yamaguti, 1953 (MPM 22557);<br />

holotype, 11 paratypes of Squamodisciis longipenis<br />

Yamaguti, 1934 (labeled as S. longiphallus) (MPM<br />

22564); and 37 voucher specimens of Calydiscoides<br />

flexuosus Yamaguti, 1953 (USNPC 89024). Host<br />

names and synonyms follow those provided by the<br />

FAO Fish Base at http://www.fao.org/waicent/faoinfo/<br />

fisher y/fi shbase/fishbase .htm.<br />

Results<br />

A total of 17 species of Diplectanidae was<br />

found on 17 species of marine fishes collected<br />

off the Kuwaiti coast. Specimens of only 7 of<br />

the 17 diplectanid species were sufficient for<br />

identification and description. Ten unidentified<br />

diplectanids and their hosts are listed in Table 1.<br />

Class Monogenoidea Bychowsky, 1937<br />

Order Dactylogyridea Bychowsky, 1937<br />

Diplectanidae Monticelli, 1903<br />

Diplectanum cazauxi Oliver and Paperna,<br />

1984<br />

(Figs. 1-8)<br />

REDESCRIPTION (measurements of specimens<br />

from Sphyraena obtusata Cuvier, 1929, follow<br />

those from Sphyraena jello Cuvier, 1829 in<br />

brackets): Diplectaninae. Body 964 (729-<br />

1,080; n = 4) [824 (608-1,070; n = 4)] long,<br />

fusiform; greatest width 170 (123-242; n = 4)<br />

[156 (97-229; n = 4)] usually in posterior trunk<br />

at level of testis. Tegument smooth. Cephalic<br />

margin tapered; 2 terminal, 2 bilateral cephalic<br />

lobes poorly developed; head organs numerous;<br />

cephalic glands numerous in cephalic area, 2 bilateral<br />

groups posterolateral to pharynx. Eyes 4;<br />

members of posterior pair slightly larger, farther<br />

apart than anterior members; 1 anterior eye occasionally<br />

absent; granules small, ovate; accessory<br />

granules absent to numerous in cephalic region.<br />

Mouth subterminal, ventral to pharynx;<br />

pharynx 52 (39-68; n = 4) [42 (32-48; n = 4)]<br />

wide, ovate to subrectangular in dorsoventral<br />

view; esophagus short or nonexistent; intestinal<br />

ceca blind. Peduncle short to elongate. Haptor<br />

81-82 (n = 2) [70 (69-72; n = 3)] long, 127<br />

(113-140; n = 2) [130 (120-137; n = 3)] wide,<br />

bilaterally lobed; squamodiscs similar, each 49<br />

(36-60; n = 6) [50 (46-61; n = 7)] long, 77<br />

Copyright © 2011, The Helminthological Society of Washington<br />

(61-88; n = 6) [72 (64-86; n = 7)] wide, with<br />

17-19 concentric rows of dumbbell-shaped rodlets,<br />

each with anterior lightly sclerotized blunt<br />

spinelet. Ventral anchor 30 (29-32; n = 11) [31<br />

(29-32; n = 6)] long, with elongate deep root,<br />

knob-like superficial root, straight shaft, moderately<br />

long point extending slightly past level<br />

of tip of superficial root; anchor base 9 (8-10;<br />

n = 3) [7-8 (n = 1)] wide. Dorsal anchor 23<br />

(22-24; n = 10) [23 (22-25; n = 8)] long, with<br />

subtriangular base, slightly curved shaft, point<br />

extending past level of tip of superficial anchor<br />

root; anchor base 7-8 (n = 6) wide. Ventral bar<br />

72 (58-85; n = 10) [66 (62-76; n = 6)] long,<br />

subrectangular, with tapered ends, ventral<br />

groove; paired dorsal bar 42 (36-46; n — 10)<br />

[40 (37-44; n = 8)] long, spatulate medially.<br />

Hooks similar; each 10 (9-11; n = 19) [10 (9-<br />

11; n = 11)] long, with protruding thumb with<br />

slightly depressed tip, delicate point, shank;<br />

hook pair 1 lying medial to anchors on short<br />

haptoral peduncles, pairs 2—4, 6 submarginal on<br />

lateral haptoral lobes, pair 5 associated with distal<br />

shaft of ventral anchor, pair 7 on dorsal surface<br />

of lateral haptoral lobe; filamentous booklet<br />

(FH) loop shank length. Male copulatory organ<br />

41 (39-44; n = 4) [36 (31-40; n = 3)] long,<br />

weakly sclerotized, C shaped, with slightly enlarged<br />

base, nipple-like termination. Accessory<br />

piece absent. Testis 261 (185-300; n = 4) [207<br />

(129-292; n = 4)] long, 92 (70-130; n = 4) [82<br />

(65—105; n = 4)] wide, pyriform; course of vas<br />

deferens not observed; seminal vesicle a simple<br />

dilation of vas deferens, lying along body midline<br />

dorsal to vagina; 2 small prostatic reservoirs<br />

immediately anterior to male copulatory organ,<br />

saccate. Ovary 42 (31-56; n = 3) [40-41 (n =<br />

1)] wide, elongate pyriform, looping right intestinal<br />

cecum, lying transversely anterior to testis;<br />

oviduct elongate; ootype ventral, a small dilated<br />

portion of female duct; uterus delicate, extending<br />

along seminal vesicle; seminal receptacle not<br />

observed; vagina nonsclerotized, aperture sinistroventral<br />

near level of male copulatory organ;<br />

vitellaria throughout trunk, except absent in regions<br />

of major reproductive organs.<br />

HOSTS AND LOCALITY: Pickhandle barracuda,<br />

Sphyraena jello Cuvier, 1829 (Sphyraenidae):<br />

Persian Gulf off Kuwait (15 October 1993). Obtuse<br />

barracuda, Sphyraena obtusata Cuvier,<br />

1829 (Sphyraenidae): Persian Gulf off Kuwait (9<br />

July 1993).<br />

PREVIOUS RECORDS: Yellowtail barracuda,


Table 1. Unidentified diplectanids infesting marine fishes off Kuwait.<br />

Host<br />

Acanthopagrus berda<br />

(Forsskal, 1775)<br />

(Sparidae)<br />

Acanthopagrus bifaxciatus<br />

(Forsskal, 1775)<br />

(Sparidae)<br />

Acanthopagrus latus<br />

(Houttnyn, 1782)<br />

(Sparidae)<br />

Diplodus noct<br />

(Valenciennes, 1830)<br />

(Sparidae)<br />

Epinephelus arcolatus<br />

(Forsskal, 1775)<br />

(Serranidae)<br />

Epinephelus tauvina<br />

(Forsskal, 1775)<br />

(Serranidae)<br />

Hemiramphus marginatus<br />

(Forsskal, 1775)<br />

(Hemiramphidae)<br />

Otolithcs argenteus<br />

(Cuvier, 1830)<br />

(Sciaenidae)<br />

Date of collection<br />

30 November 1996<br />

10 May 1995<br />

10 October 1995<br />

28 March 1995<br />

23 March 1996<br />

15 October 1994<br />

29 July 1993<br />

16 June 1993<br />

15 October 1994<br />

15 June 1993<br />

10 March 1994<br />

8 May 1995<br />

18 October 1995<br />

5 April 1996<br />

15 October 1993<br />

Sphyraena flavicauda Riippell, 1838 (Sphyraenidae):<br />

Gulf of Aqaba (Golfe D'Aquaba [sic]),<br />

Gulf of Suez (Egypt), Indian Ocean off Malindi<br />

(Kenya) (all Oliver and Paperna, 1984).<br />

SPECIMENS STUDIED: 12 voucher specimens<br />

from S. jello, USNPC 89010, HWML 15023; 8<br />

voucher specimens from S. obtusata, USNPC<br />

89009.<br />

REMARKS: Diplectanum cazauxi is known<br />

only from species of barracuda (Sphyraenidae).<br />

Our report of this species on 5. jello and S. obtusata<br />

from the Persian Gulf represents new host<br />

and geographic records. The known geographic<br />

distribution of D. cazauxi currently includes the<br />

western Indian Ocean and adjacent regions including<br />

the northern gulfs of the Red Sea and<br />

the Persian Gulf.<br />

The original description of D. cazauxi is<br />

based on morphometrics of the squamodisc and<br />

sclerotized haptoral and copulatory structures.<br />

Although Oliver and Paperna (1984) mentioned<br />

that the ovary loops the right intestinal cecum,<br />

a symplesiomorphic feature for all members of<br />

the Diplectanidae, other details of the internal<br />

anatomy were not considered. Our redescription<br />

KRITSKY ET AL.—DIPLECTANIDS FROM KUWAIT 147<br />

Parasite<br />

Lamellodiscus sp. 1<br />

Lamellodiscus sp. 2<br />

Lamellodiscus sp. 1<br />

Lamellodiscus sp. 3<br />

Lamellodiscus sp. 4<br />

Lamellodiscus sp. 5<br />

Pseudorhabdosynochus sp. 1<br />

Pseudorhabdosynochus sp. 2<br />

Lamellodiscus sp. 6<br />

Diplectanum sp. 1<br />

Diplectanurn sp. 2<br />

USNPC no.<br />

89011<br />

89012<br />

89013<br />

89014<br />

89015<br />

89016<br />

89017<br />

89018<br />

89019<br />

89030<br />

89029<br />

89031<br />

89034<br />

89033<br />

89035<br />

89032<br />

adds information on soft-tissue features of the<br />

reproductive, digestive, and nervous systems.<br />

The morphometrics of the haptoral sclerites<br />

and squamodisc in our specimens are in general<br />

agreement with those reported by Oliver and<br />

Paperna (1984) in the original description of D.<br />

cazauxi. Mounting media (Gray and Wess' medium,<br />

Malmberg's medium, and Hoyer's medium)<br />

commonly used to visualize the sclerites of<br />

monogenoideans apply pressure on the specimen.<br />

In D. cazauxi, this pressure results in significant<br />

distortion of the lightly sclerotized male<br />

copulatory organ. The copulatory organs of D.<br />

cazauxi shown in Figure 11 of Oliver and Paperna<br />

(1984) are clearly distorted, as were our<br />

specimens mounted in Gray and Wess' medium.<br />

Such artifacts are minimized when specimens<br />

are mounted in Canada balsam, which does not<br />

result in significant coverslip pressure on the<br />

specimen (compare Fig. 4 with Fig. 11 of Oliver<br />

and Paperna, 1984).<br />

The copulatory complex, dorsal anchor, haptoral<br />

bars, and squamodisc of Diplectanum cazauxi<br />

closely resemble those of Laterocaecum<br />

pearsoni Young, 1969, suggesting that these<br />

Copyright © 2011, The Helminthological Society of Washington


148 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Figures 1-8. Diplectanum cazauxi Oliver and Paperna, 1984. 1. Whole mount (composite, ventral;<br />

dorsal squamodisc not shown). 2. Ventral anchor. 3. Dorsal anchor. 4. Copulatory complex. 5. Dorsal bar.<br />

6. Hook. 7. Ventral bar. 8. Ventral view of haptor showing ventral squamodisc and positions of hook<br />

pairs. All figures are drawn to the 25-jjim scale, except Figures 1 and 8 (200-jjim and 50-u.m scales,<br />

respectively).<br />

species likely share a common evolutionary history.<br />

Laterocaecum was proposed by Young<br />

(1969) for a diplectanid collected from the obtuse<br />

barracuda, S. obtusata, from Moreton Bay,<br />

Queensland, Australia. Young (1969) differentiated<br />

the genus from other diplectanid genera<br />

Copyright © 2011, The Helminthological Society of Washington<br />

by species possessing lateral diverticula of the<br />

intestinal ceca (lateral diverticula absent in all<br />

other species of Diplectanidae) and 12 (6 pairs)<br />

hooks in the adult. If D. cazauxi actually shares<br />

a phylogenetic history with L. pearsoni as suggested<br />

by their similar morphology and host


preferences, separation of Laterocaecum from<br />

Diplectanum may not be justified, and the 2<br />

unique characters presented by L. pearsoni may<br />

represent secondarily derived features within Diplectanum.<br />

We do not formally propose synonymy<br />

of the 2 genera at this time, however, because<br />

hypotheses on phylogenetic relationships<br />

within the Diplectanidae are lacking and Diplectanum<br />

may represent a paraphyletic group (see<br />

"Discussion"). Diplectanum cazauxi differs<br />

from L. pearsoni by having a knob-like superficial<br />

root on the ventral anchor (root elongate<br />

in L. pearsoni) and by possessing 7 pairs of<br />

hooks in the adult (6 pairs in L. pearsoni).<br />

Diplectanum sillagonum Tripathi, 1957<br />

(Figs. 9-15)<br />

REDESCRIPTION (Tripathi's [1957] original<br />

measurements and counts are in brackets following<br />

respective parameters of specimens from the<br />

Persian Gulf): Diplectaninae. Body 755 (694-<br />

815; n = 4) [623-1,058] long, fusiform, somewhat<br />

flattened dorsoventrally; greatest width 131<br />

(110-153; n = 4) [114-144] usually in anterior<br />

trunk near level of copulatory organ. Tegument<br />

smooth. Cephalic margin tapered; 2 terminal, 2<br />

bilateral cephalic lobes poorly developed; subspherical<br />

ventral pouch lying anterior to pharynx,<br />

opening to exterior via simple midventral<br />

pore. Head organs numerous; distributed in 3<br />

poorly defined groups; anterior posterior groups<br />

associated with respective cephalic lobes. Cephalic<br />

glands lateral to pharynx, extending posteriorly<br />

past level of esophageal bifurcation.<br />

Eyes 4; members of posterior pair larger, closer<br />

together than anterior members; granules small,<br />

ovate; accessory granules numerous, distributed<br />

throughout cephalic, anterior trunk regions.<br />

Mouth subterminal, ventral to pharynx; pharynx<br />

47 (40-53; n = 4) [41-49] wide, subspherical;<br />

esophagus short or absent; intestinal ceca blind.<br />

Peduncle short, broad. Haptor 124 (113-137;<br />

n = 4) [57] long, 159 (150-170; n = 4) [133-<br />

152] wide, bilaterally lobed; squamodiscs similar,<br />

each 73 (62-83; n = 12) [57-76] in diameter,<br />

subcircular, with 13—15 [11—15] concentric<br />

rows of dumbbell-shaped rodlets, each with anterior<br />

lightly sclerotized blunt spinelet. Ventral<br />

anchor 44 (38-50; n = 14) [49-53] long, with<br />

elongate roots (deep root longest), straight shaft,<br />

recurved point extending slightly past level of<br />

tip of superficial anchor root; anchor base 14<br />

KRITSKY ET AL.—DIPLECTANIDS FROM KUWAIT 149<br />

(11-16; n = 8) wide. Dorsal anchor 40 (38-44;<br />

n = 13) [41-49] long, with subtriangular base,<br />

slightly curved shaft, recurved point extending<br />

past level of tip of superficial anchor root; anchor<br />

base 12 (10-14; n = 7) wide. Ventral bar<br />

74 (<strong>67</strong>-86; n = 10) [60-72] long, with tapered<br />

ends, ventral groove; median anterior constriction.<br />

Paired dorsal bar 69 (63-75; n = 11) [57-<br />

64] long, medial end expanded, bilobed. Hooks<br />

similar; each 12 (11—13; n = 29) long, with protruding<br />

thumb with slightly depressed tip, delicate<br />

point, slender shank; hook pair 1 at level of<br />

tips of ventral bar, medial to anchors; pairs 2—4,<br />

6, 7 submarginal in lateral haptoral lobes; pair 5<br />

associated with distal ventral anchor shaft; FH<br />

loop shank length. Male copulatory organ 34<br />

(30-39; n = 6) [41-45] long, a sigmoid tube<br />

originating from ring-like sclerotized base, with<br />

fine recurved tip. Accessory piece variable,<br />

comprising 2 articulated subunits, 1 subunit with<br />

bilobed proximal end articulating to other subunit.<br />

Testis 70 (69-71; n = 2) [38-53 X 76-<br />

152] in diameter, subspherical; course of vas deferens<br />

not observed; seminal vesicle a simple<br />

elongate dilation of vas deferens, lying along<br />

body midline dorsal to seminal receptacle; prostatic<br />

reservoir saccate, posterior to male copulatory<br />

organ, frequently containing granules<br />

only at anterior end. Ovary 57 (42-71; n = 2)<br />

[38 X 57] wide, elongate pyriform, looping right<br />

intestinal cecum, lying transversely anterior to<br />

testis; oviduct elongate; ootype, uterus not observed;<br />

seminal receptacle ovate, originating<br />

from short tubular vagina; vagina with small<br />

bead-like sclerotization having cupped proximal<br />

end; vaginal aperture sinistral; vitellaria throughout<br />

trunk, except absent in regions of major reproductive<br />

organs.<br />

HOSTS AND LOCALITY: Silver sillago, Sillago<br />

sihama (Forsskal, 1775) (Sillaginidae): Persian<br />

Gulf off Kuwait (31 December 1993, 18 April<br />

1996).<br />

PREVIOUS RECORDS: Sillago sihama: Chandipore,<br />

Chilka Lake, Puri, all Bay of Bengal,<br />

India (Tripathi, 1957). Sillago sihama: Burdekin<br />

River, Duyfken Point, Point Samson, and Darwin,<br />

Australia; Phuket and Bang Saen, Thailand;<br />

Gendering and Kula Lumpur, Malaysia; Bali, Indonesia;<br />

Aberdeen market and Sai Kung, Hong<br />

Kong; Ring Ring, Kapa Kapa, and Sinapa, Paupua<br />

New Guinea; and Madras, India (all Hayward,<br />

1996). Slender sillago, Sillago attenuata<br />

McKay, 1985: Ras Lanura, Saudi Arabia (Hay-<br />

Copyright © 2011, The Helminthological Society of Washington


150 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

9<br />

Figures 9-15. Diplectanum sillagonum Tripathi, 1957. 9. Whole mount (composite, body ventral, haptor<br />

dorsal), showing position of hook pairs. 10. Hook. 11. Copulatory complex. 12. Ventral bar. 13. Dorsal<br />

bar. 14. Ventral anchor. 15. Dorsal anchor. All figures are drawn to the 25-u.m scale, except Figure 9<br />

(200-|xm scale).<br />

10<br />

Copyright © 2011, The Helminthological Society of Washington<br />

15


ward, 1996). Vincent's sillago, Sillago vincenti<br />

McKay, 1980: Kavanad, Kerala, India (Hayward,<br />

1996).<br />

SPECIMENS STUDIED: 14 voucher specimens,<br />

USNPC 89007, 89008, HWML 15022.<br />

REMARKS: Diplectanum sillagonum was described<br />

by Tripathi (1957) from the gills of<br />

S. sihama from western coastal localities on the<br />

Bay of Bengal, India. His description of this species<br />

is of marginal value for species determination.<br />

Nonetheless, the original drawings of the<br />

copulatory complex, anchors, bars, and whole<br />

mount, while diagrammatic, strongly suggest<br />

conspecificity with our collection from the Persian<br />

Gulf. Persian Gulf specimens were obtained<br />

from the same host species as that of the type<br />

series, and respective measurements of specimens<br />

from the Persian Gulf and India are comparable.<br />

However, the types of D. sillagonum<br />

were not available for confirmation. General<br />

morphology of the sclerotized haptoral structures<br />

and copulatory complex generally corresponds<br />

to figures of this species offered by Hayward<br />

(1996). However, Hay ward (1996) did not<br />

mention the presence of the midventral pouch<br />

located anterior to the pharynx in his redescription<br />

of the species.<br />

Lepidotrema kuwaitensis sp. n.<br />

(Figs. 16-23)<br />

DESCRIPTION: Diplectaninae. Body 504<br />

(452—603; n — 8) long, fusiform; greatest width<br />

105 (90-121; n = 9) near body midlength. Tegument<br />

smooth. Cephalic margin tapered; 2 terminal,<br />

2 bilateral cephalic lobes poorly developed;<br />

3 bilateral pairs of head organs with anterior,<br />

posterior pairs associated with respective<br />

cephalic lobes; cephalic glands not observed.<br />

Eyes 4, equidistant; members of posterior pair<br />

larger than anterior members; anterior eyes frequently<br />

absent, 1 or both posterior eyes occasionally<br />

dissociated; granules small, ovate, numerous<br />

accessory granules at eye level. Mouth<br />

subterminal, ventral to pharynx; pharynx 23<br />

(19-26; n = 9) wide, ovate to subspherical;<br />

esophagus short to absent; intestinal ceca blind.<br />

Peduncle short to elongate. Haptor 83 (65-100;<br />

n = 9) long, 139 (124-151; n = 9) wide, bilaterally<br />

lobed. Squamodiscs similar, each 37 (26-<br />

43; n = 4) long, 40 (27-48; n = 6) wide, subcircular,<br />

with 8-10 concentric rows of dumbbellshaped<br />

rodlets becoming progressively more<br />

KRITSKY ET AL.—DIPLECTANIDS FROM KUWAIT 151<br />

delicate in posterior rows; 2-4 rows (layers) of<br />

elongate delicate spinelets wrap around posterior<br />

margin of both squamodiscs, spinelets frequently<br />

absent. Ventral anchor 42 (39-45; n = 25)<br />

long, with elongate roots (deep root longest),<br />

evenly curved shaft with terminal indentation at<br />

articulation with recurved point; point extending<br />

slightly past level of tip of superficial anchor<br />

root; anchor base 9 (7-11; n = 13) wide. Dorsal<br />

anchor 37 (32-40; n = 23) long, with narrow<br />

base, long deep root, curved shaft, point extending<br />

past level of tip of superficial root of anchor<br />

base; anchor base 7 (6-9; n = 11) wide. Ventral<br />

bar 59 (52-66; n = 21) long, with tapered ends,<br />

ventral groove; paired dorsal bar 55 (47—60; n<br />

= 23) long, spatulate, with posteromedial spine.<br />

Hook 10 (9-12; n = 37) long, with protruding<br />

thumb with slightly depressed end, delicate<br />

point, shank dilated slightly in some specimens.<br />

Hook pair 1 lying medial to haptoral lobes, posterior<br />

to bars; pairs 2-4, 7 in lateral haptoral<br />

lobes; pair 5 associated with shaft of ventral anchor;<br />

pair 6 at level of or just anterior to dorsal<br />

bar. FH loop shank length. Male copulatory organ<br />

68 (60-74; n =11) long, a sigmoid tube<br />

with wall of varying thickness along length,<br />

acute tip. Accessory piece absent. Testis subspherical,<br />

54 (42-65; n = 9) in diameter; course<br />

of vas deferens not observed; seminal vesicle a<br />

simple dilation of vas deferens, lying along body<br />

midline dorsal to ootype; prostatic reservoirs 3,<br />

saccate; anterior prostatic vesicles bilateral to<br />

male copulatory organ, with prostatic ducts<br />

fused prior to entering base of male copulatory<br />

organ via common duct; posterior reservoir caudal<br />

to male copulatory organ, apparently emptying<br />

independently into base of male copulatory<br />

organ. Ovary 23 (19—25; n = 3) wide, pyriform,<br />

anterodorsal to testis, looping right intestinal<br />

cecum; oviduct elongate; ootype ventral, a<br />

small dilated portion of female duct; uterus delicate,<br />

extending anteriorly to left of prostatic reservoirs;<br />

seminal receptacle not observed; vaginal<br />

aperture sinistroventral, with circular muscular<br />

rim; vagina funnel-shaped, narrowing to<br />

short tube; vagina with proximally thickened<br />

walls; vitellaria throughout trunk, except absent<br />

in regions of major reproductive organs. Egg<br />

83-84 (n = 1) long, 56-57 (n = 1) wide, ovate,<br />

with short proximal filament.<br />

TYPE HOST: Small-scaled terapon, Terapon<br />

puta (Cuvier, 1829) (Terapontidae).<br />

Copyright © 2011, The Helminthological Society of Washington


152 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Figures 16—23. Lepidotrema kuwaitensis sp. n. 16. Whole mount (composite, ventral; dorsal squamodisc<br />

not shown), showing positions of hook pairs. 17. Copulatory complex. 18. Hooks. 19. Enlargement of worm<br />

at level of reproductive organs (composite, ventral). 20. Dorsal bar. 21. Ventral bar. 22. Ventral anchor.<br />

23. Dorsal anchor. All figures are drawn to the 25-u.m scale, except Figures 16 and 19 (100-u.m and 50ujm<br />

scales, respectively).<br />

TYPE LOCALITY: Persian Gulf off Kuwait (9<br />

July 1993, 15 October 1993, 26 March 1996).<br />

INFECTION SITE: Gills.<br />

DEPOSITED SPECIMENS: Holotype, USNPC<br />

89020; 25 paratypes, USNPC 89021, 89022,<br />

89023, HWML 15025.<br />

ETYMOLOGY: This species is named for the<br />

country of Kuwait.<br />

REMARKS: The primary distinguishing feature<br />

of Lepidotrema Johnston and Tiegs, 1922, is the<br />

presence of groups of elongate spinelets forming<br />

fan-like structures on the posterior portions of the<br />

Copyright © 2011, The Helminthological Society of Washington<br />

squamodiscs (Oliver, 1987). The genus currently<br />

includes 6 species, all from freshwater teraponids<br />

in Australia: Lepidotrema therapon Johnston and<br />

Tiegs, 1922; L. angusta; L. bidyana; Lepidotrema<br />

fidiginosum Johnston and Tiegs, 1922; Lepidotrema<br />

simplex Johnson and Tiegs, 1922; and L. tentie.<br />

Our finding of L. kuwaitensis on T. puta (Teraponidae)<br />

in the Persian Gulf is the first report of<br />

a member of Lepidotrema from a marine host.<br />

Existing descriptions of the 6 freshwater species<br />

are of marginal value for comparison with L. kuwaitensis,<br />

and most species require redescription


(see Johnston and Tiegs, 1922; Murray, 1931;<br />

Young, 1969).<br />

In L. kuwaitensis, the posterior spinelets are<br />

delicate (or frequently absent, an apparent artifact<br />

resulting from deterioration of the specimen before<br />

fixation) and resemble those of L. angusta<br />

as depicted by Young (1969). These species are<br />

easily separated by the comparative morphology<br />

of the copulatory complexes (sigmoid in L. kuwaitensis;<br />

coiled with about 1 ring in L. angusta).<br />

Examination of the types of Diplectanum longipenis<br />

(Yamaguti, 1934) Yamaguti, 1963<br />

(=Squamodiscus longipenis Yamaguti, 1934),<br />

confirmed that L. kuwaitensis shares many features<br />

(general morphology and arrangement of<br />

the sclerotized haptoral and copulatory sclerites<br />

and internal reproductive organs) with this species<br />

and may be more closely aligned to it than<br />

to those from fresh water. While staining procedures<br />

used by Yamaguti (1934) did not allow<br />

us to see spinelets near the posterior margin of<br />

the squamodisc in D. longipenis, similarities in<br />

the morphology of sclerotized structures and the<br />

general organization of the reproductive organs<br />

suggest that the 2 species are congeneric. Thus,<br />

we propose the transfer of D. longipenis to Lepidotrema<br />

as L. longipenis (Yamaguti, 1934)<br />

comb. n. Squamodiscus Yamaguti, 1934, is removed<br />

from synonymy with Diplectanum and<br />

becomes a junior subjective synonym of Lepidotrema.<br />

Lepidotrema kuwaitensis differs from<br />

L. longipenis by having delicate anchors (base<br />

of dorsal anchor in L. kuwaitensis narrow; broad<br />

in D. longipenis) and by the number of rodlet<br />

rows in the squamodisc (8—10 rows in L. kuwaitensis;<br />

18-21 in D. longipenis).<br />

Pseudolamellodiscus sphyraenae Yamaguti,<br />

1953<br />

(Figs. 24-34)<br />

REDESCRIPTION: Diplectaninae. Body 1196<br />

(1020-1354; n = 17) long, flattened dorsoventrally;<br />

greatest width 244 (196-289; n = 16)<br />

usually in anterior trunk. Trunk with anterior<br />

dextroventral sclerite, posterior dextroventral<br />

sclerite, sinistroventral spined pit. Anterior dextroventral<br />

sclerite 58 (48-72; n = 26) long, with<br />

lobulate base, rod-shaped distal end protruding<br />

from small ventral pore, spined; number of<br />

spines variable. Posterior dextroventral sclerite<br />

37 (32-45; n = 27) long, spatulate, with incised<br />

distal margin; sinistroventral pit blind, with 4—6<br />

KRITSKY ET AL.—DIPLECTANIDS FROM KUWAIT 153<br />

spines, opening ventrally via small aperture<br />

through tegument; tips of spines usually protruding<br />

through pore. Tegument smooth. Cephalic<br />

margin tapered; cephalic lobes poorly developed;<br />

head organs numerous along anterolateral<br />

margins of cephalic area; cephalic glands<br />

posterolateral to pharynx. Eyes 4; members of<br />

posterior pair larger, slightly farther apart than<br />

anterior members; 1 member of each pair occasionally<br />

absent; granules small, irregular; accessory<br />

granules uncommon in cephalic region.<br />

Mouth subterminal, ventral to anterior portion of<br />

pharynx; pharynx 61 (53-70; n = 19) wide,<br />

elongate, ovate; esophagus short to nonexistent;<br />

intestinal ceca blind. Peduncle broad. Haptor<br />

337 (260-421; n = 18) wide, 114 (93-148;<br />

n = 18) long, bilaterally lobed; squamodiscs<br />

similar, each 61 (47-70; n = 17) long, 249<br />

(200-310; n = 17) wide, with approximately 45<br />

longitudinal parallel rows of dumbbell-shaped<br />

spines in anterior portion of squamodisc; posterior<br />

portion with numerous spine-like scales.<br />

Ventral anchor 41 (36-44; n = 25) long, with<br />

elongate deep root, knob-like superficial root,<br />

slightly curved shaft, recurved point extending<br />

past level of tip of superficial anchor root; anchor<br />

base 11 (10—12; n = 3) wide. Dorsal anchor<br />

33 (31-36; n = 31) long, with short deep<br />

root, triangular superficial root perpendicular to<br />

anchor base, curved shaft, point extending past<br />

level of tip of superficial root of anchor base;<br />

anchor base 9 (8-11; n = 6) wide. Ventral bar<br />

268 (218-328; n = 22) long, narrowed medially,<br />

ends tapered, recurved anteriorly; ventral groove<br />

present. Paired dorsal bar 69 (59-87; n = 25)<br />

long, club-shaped. Hooks similar; each 10-11 (n<br />

= 26) long, with protruding depressed thumb,<br />

delicate point, shank. Hook pair 1 submarginal,<br />

lying posterior to bars near base of haptoral<br />

lobes; pairs 2-7 located on lateral haptoral<br />

lobes; FH loop shank length. Male copulatory<br />

organ 33 (31-35; n = 9) long, with large base,<br />

bent shaft, acute bent tip, subbasal pointed projection.<br />

Accessory piece absent. Common genital<br />

pore absent; male genital pore lying ventrally<br />

to left of body midline slightly posterior to male<br />

copulatory organ; uterine pore ventral, slightly<br />

posterior to level of male genital pore, somewhat<br />

dextral to body midline. Testis 165 (144-184; n<br />

= 15) long, 81 (59-98; n = 16) wide, ovate;<br />

course of vas deferens not observed; 2 seminal<br />

vesicles simple dilations of vas deferens; proximal<br />

vesicle elongate, fusiform, lying along mid-<br />

Copyright © 2011, The Helminthological Society of Washington


154 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

5 1 34<br />

Figures 24-34. Pseudolamellodiscus sphyraenae Yamaguti, 1953. 24. Whole mount (composite, ventral;<br />

dorsal squamodisc not shown), showing positions of hook pairs. 25. Anterior dextroventral sclerite. 26.<br />

Copulatory complex. 27. Posterior dextroventral sclerite. 28. Posterodorsal accessory sclerite (lateral view).<br />

29. Sinistroventral spinous cavity (lateral view). 30. Hook. 31. Dorsal bar. 32. Dorsal anchor. 33. Ventral<br />

anchor. 34. Ventral bar. All figures are drawn to the 25-u.m scale, except Figures 24 and 34 (200-jim and<br />

50-fjim scales, respectively).<br />

line of body posterior to male copulatory organ;<br />

distal vesicle anterior to male copulatory organ,<br />

short, pyriform; prostatic reservoir saccate, anterior<br />

to male copulatory organ. Ovary 81 (62-<br />

107; n = 16) wide, forming lobed cap on anterior<br />

margin of testis, with proximal sinistral loop<br />

before extending around right intestinal cecum;<br />

oviduct narrowing to small tube before joining<br />

Copyright © 2011, The Helminthological Society of Washington<br />

slightly expanded ootype; uterus delicate, extending<br />

to right of body midline; vaginal aperture<br />

sinistroventral; vagina tubular, frequently<br />

containing apparent spermatophore, joining<br />

small seminal receptacle lying to left of ootype;<br />

vitellaria throughout trunk, except absent in regions<br />

of reproductive organs.<br />

HOST AND LOCALITY: Yellowstrip barracuda,


Sphyraena chiysotaenia Klunzinger, 1884<br />

(Sphyraenidae): Persian Gulf off Kuwait (16 October<br />

1996).<br />

PREVIOUS RECORDS: Sphyraena sp.: Macassar,<br />

Celebes (Yamaguti, 1953). Great barracuda,<br />

Sphyraena barracuda (Walbaum, 1972): Nosy<br />

Be, Madagascar (Rakotofiringa and Maillard,<br />

1979).<br />

SPECIMENS STUDIED: 34 voucher specimens,<br />

USNPC 89028, HWML 15020.<br />

REMARKS: While Yamaguti (1953) did not<br />

adequately describe the haptoral sclerites, male<br />

copulatory organ, and trunk sclerites of Pseudolamellodiscus<br />

sphyraenae, our examination of<br />

the holotype and paratypes of this species confirmed<br />

that our specimens were conspecific with<br />

P. sphyraenea. In the account of P. sphyraenea<br />

from Madagascar by Rakotofiringa and Maillard<br />

(1979), the morphology of the haptoral and<br />

trunk sclerites were also not presented, but their<br />

figure of the trunk region, which includes a<br />

small drawing of the male copulatory organ,<br />

clearly supports their identification. However,<br />

both Yamaguti (1953) and Rakotofiringa and<br />

Maillard (1979) confused the vagina with the<br />

uterus. This error is supported by some of our<br />

specimens that contained a spermatophore in the<br />

tube that these authors described as the "uterus"<br />

and a developing egg in the tube they labeled<br />

"vagina."<br />

The slide containing the types of P. sphyraenea<br />

includes the holotype, 23 paratypes, and<br />

several fragments of specimens. Included in the<br />

23 paratypes are 2 specimens clearly of an undescribed<br />

Pseudolamellodiscus species, characterized<br />

by having 1 large ventral trunk sclerite<br />

with a bifurcated, foliated proximal end.<br />

Lamellodiscus furcillatus sp. n.<br />

(Figs. 35-42)<br />

DESCRIPTION: Lamellodiscinae. Body 1,092<br />

(924-1,294; n = 4), long, fusiform; greatest<br />

width 207 (183-226; n = 4), usually in posterior<br />

trunk near level of testis. Tegument smooth. Cephalic<br />

margin tapered; 2 terminal, 2 bilateral cephalic<br />

lobes poorly developed; 3 bilateral pairs<br />

of head organs with anterior, posterior pairs associated<br />

with respective cephalic lobes; cephalic<br />

glands posterolateral to pharynx. Eyes 4; equidistant;<br />

members of posterior pair larger than<br />

anterior members; anterior eyes occasionally absent;<br />

granules ovate, variable in size; accessory<br />

KRITSKY ET AL.—DIPLECTANIDS FROM KUWAIT 155<br />

granules common in cephalic region. Mouth<br />

subterminal, ventral to pharynx; pharynx 54<br />

(46-60, n = 4) wide, ovate to subspherical; bilateral<br />

pair of prepharyngeal (buccal) glands anterior<br />

to pharynx; esophagus short to nonexistent;<br />

intestinal ceca blind. Peduncle broad. Haptor<br />

163 (148-180; n = 4) wide, 111 (104-117;<br />

n = 4) long, bilaterally lobed; lobes short. Lamellodiscs<br />

similar; each 58 (52-62; n = 4) long,<br />

42 (40-46; n — 4) wide, ovate, with 10 lamellar<br />

rings; anterior (deep) lamella forming complete<br />

ring; intermediate lamellae superficially incomplete,<br />

medially indented; posterior (superficial)<br />

lamella indented, complete. Ventral anchor 58<br />

(54-61; n — 8) long, with elongate deep root,<br />

short depressed superficial root, evenly curved<br />

shaft, recurved point; point extending slightly<br />

past level of tip of superficial anchor root; anchor<br />

base 11-12 (n = 2) wide. Dorsal anchor 48<br />

(45-52; n = 8) long, with elongate deep root,<br />

erect knob-like superficial root, evenly curved<br />

shaft, nonrecurved point; anchor base 14-15<br />

(n = 2) wide. Ventral bar 76 (70-82; n = 8)<br />

long, plate-like, with ends constricted subterminally,<br />

ventral groove. Paired dorsal bar 62<br />

(56-68; n = 8) long, morphologically complex,<br />

broad. Hooks similar; each 12 (11-13; n = 7)<br />

long, with protruding slightly depressed thumb,<br />

delicate point, shank; FH loop shank length.<br />

Hook pair 1 submedial at level of posterior margin<br />

of ventral bar; pairs 2-4 submarginal in lateral<br />

haptoral lobes; pair 5 associated with ventral<br />

anchor shafts; pairs 6, 7 dorsal at level of tip of<br />

deep root of ventral anchor. Male copulatory organ<br />

60 (58-65; n = 7) long, a sigmoid tube with<br />

acute recurved tip. Accessory piece 53 (49—58;<br />

n = 4) long, with subterminal elongate branch.<br />

Testis subspherical, 111 (110-113; n = 2) in diameter,<br />

course of vas deferens not observed;<br />

seminal vesicle a simple dilation of vas deferens,<br />

lying to left of body midline dorsal to seminal<br />

receptacle; prostatic reservoir saccate, lying anterior<br />

to copulatory complex. Ovary 65 (64—<strong>67</strong>;<br />

n = 2) wide, elongate pyriform, diagonal, looping<br />

right intestinal cecum, overlapping testis;<br />

oviduct elongate; ootype ventral, a small dilated<br />

portion of female duct; uterus delicate, seminal<br />

receptacle small. Vaginal aperture sinistral; vagina<br />

short, frequently containing apparent subspherical<br />

spermatophore. Vitellaria throughout<br />

trunk, except absent in regions of reproductive<br />

organs.<br />

Copyright © 2011, The Helminthological Society of Washington


156 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Figures 35-42. Lamellodiscus furcillatus sp. n. 35. Whole mount (composite, dorsal; ventral lamellodisc<br />

not shown). 36. Dorsal bar. 37. Ventral bar. 38. Copulatory complex. 39. Ventral anchor. 40.<br />

Hook. 41. Dorsal anchor. 42. Dorsal view of haptor showing dorsal lamellodisc and positions of hook<br />

pairs. All figures are drawn to the 25-fjim scale, except Figures 35 and 42 (500-u,m and 50-fj.m scales,<br />

respectively).<br />

TYPE HOST: Red Sea seabream, Diplodus<br />

noct (Valenciennes, 1830) (Sparidae).<br />

TYPE LOCALITY: Persian Gulf off Kuwait (27<br />

October 1995, 23 March 1996).<br />

INFECTION SITE: Gills.<br />

DEPOSITED SPECIMENS: Holotype, USNPC<br />

89036; 7 paratypes, USNPC 89037, HWML<br />

15024.<br />

ETYMOLOGY: The specific name is from Lat-<br />

Copyright © 2011, The Helminthological Society of Washington<br />

in (furcillatus = a small fork) and refers to the<br />

accessory piece of the copulatory complex.<br />

REMARKS: Lamellodiscus furcillatus sp. n.<br />

resembles Lamellodiscus baeri Oliver, 1974, in<br />

the morphology of the paired dorsal bars and<br />

general morphology of the copulatory complex.<br />

Oliver's (1974) description of L. baeri from the<br />

common seabream, Pagrus pagrus (Linnaeus,<br />

1758), Sparidae, is brief and does not include


figures of the internal anatomy (whole mount),<br />

hooks, or lamellodisc. However, L. furcillatus is<br />

easily differentiated from L. baeri by the presence<br />

of a nonrecurved point of the dorsal anchor<br />

(point of dorsal anchor recurved in L. baeri).<br />

Calydiscoides flexuosus (Yamaguti, 1953)<br />

Young, 1969<br />

(Figs. 43-52)<br />

REDESCRIPTION (Table 2 for measurements):<br />

Lamellodiscinae. Body long, fusiform; greatest<br />

width usually at level of testis. Tegument<br />

smooth. Cephalic margin tapered; 2 terminal, 2<br />

bilateral cephalic lobes poorly developed; 3 bilateral<br />

groups of head organs with anterior, posterior<br />

groups associated with respective cephalic<br />

lobes; cephalic glands posterolateral to pharynx.<br />

Eyes 4; members of posterior pair larger, usually<br />

closer together than anterior members; 1 member<br />

of anterior pair occasionally absent; granules<br />

usually ovate, variable in size; accessory granules<br />

common in cephalic region. Mouth subterminal,<br />

ventral to pharynx; pharynx ovate to subspherical;<br />

esophagus short; intestinal ceca blind.<br />

Peduncle broad. Haptor bilaterally lobed; lamellodiscs<br />

similar, each with 10 "telescoping" lamellae,<br />

posterior lamellae incomplete forming<br />

posterior superficial opening. Ventral anchor<br />

with elongate roots (superficial root longest)<br />

usually overlying one another (Fig. 51), evenly<br />

curved shaft, point recurved, not reaching level<br />

of tip of superficial anchor root. Dorsal anchor<br />

with short deep root, triangular superficial root,<br />

curved shaft, point extending past level of tip of<br />

superficial root. Ventral bar with tapered ends<br />

directed anterolaterally, ventral groove. Paired<br />

dorsal bar with bilobed medial end. Hooks similar;<br />

each with protruding thumb with slightly<br />

depressed end, delicate point, shank; hook pair<br />

1 lying medial to ventral anchor at level of anterior<br />

margin of ventral bar, pairs 2 (anterior), 3<br />

lateral to ventral lamellodisc, pairs 4, 6 submarginal<br />

in haptoral lobe, pair 5 associated with<br />

ventral anchor shaft, pair 7 lateral to dorsal lamellodisc;<br />

FH loop nearly shank length. Male<br />

copulatory organ, accessory piece nonarticulated.<br />

Male copulatory organ C-shaped, variably<br />

sclerotized, with acute termination, 2 subterminal<br />

branches embedded in wall of genital atrium<br />

present or absent. Accessory piece variable, flattened.<br />

Testis ovate; course of vas deferens in relation<br />

to gut not observed; vas deferens tortuous<br />

KRITSKY ET AL.—DIPLECTANIDS FROM KUWAIT 157<br />

(Fig. 46), with anterior loop, expanded to form<br />

seminal vesicle; prostatic reservoir not observed.<br />

Ovary elongate pyriform, looping right<br />

intestinal cecum, lying transversely to diagonally<br />

anterodorsal to testis; oviduct elongate; ootype<br />

an expanded portion of female duct, surrounded<br />

by numerous glands; uterus with thick<br />

wall, ventral to proximal portion of vas deferens,<br />

extending dorsal to anterior loop of vas deferens.<br />

Vaginal aperture sinistroventral; vagina variable,<br />

funnel-shaped, narrowing to short tortuous tube;<br />

seminal receptacle absent, or represented by<br />

small expansion of vaginal duct prior to emptying<br />

into female duct; vaginal funnel with sclerotized<br />

clasp-like wall. Vitellaria coextensive<br />

with gut, absent in regions of reproductive organs.<br />

SYNONYMS: Lamellodiscus flexuosus Yamaguti,<br />

1953; Lamellospina Indiana Karyakarte<br />

and Das, 1978; Calydiscoides indicus Venkatanarsaiah<br />

and Kulkarni, 1980.<br />

HOST AND LOCALITIES: Notchedfin threadfin<br />

bream, Nemipterus peronii (Valenciennes, 1830)<br />

(originally identified as N. tolu [Valenciennes,<br />

1830]) and Delagoa threadfin bream, Nemipterus<br />

bipunctatus (Valenciennes, 1830) (originally<br />

identified as N. delagoae Smith, 1941) (Nemipteridae):<br />

Persian Gulf off Kuwait (31 December<br />

1993, 10 January 1994, respectively). Japanese<br />

threadfin bream, Nemipterus japonicus<br />

(Bloch, 1791) (Nemipteridae): Port of Okha,<br />

Gujarat, India.<br />

PREVIOUS RECORDS: Ornate threadfin bream,<br />

Nemipterus hexodon (Quoy and Gaimard, 1824)<br />

(originally identified as Synagris taeniopterus<br />

(Valenciennes, 1830): Macassar, Celebes (Yamaguti,<br />

1953). Nemipterus japonicus: Ratnagiri,<br />

west coast, Maharashtra, India (Karyakarte and<br />

Das, 1978); Kakinada, Bay of Bengal, India<br />

(Venkatanarsaiah and Kulkarni, 1980).<br />

SPECIMENS STUDIED: 18 voucher specimens<br />

from N. peronii, USNPC 89025, HWML<br />

15019; 14 voucher specimens from N. bipunctatus,<br />

USNPC 89026; 37 voucher specimens<br />

from N. japonicus, USNPC 89024.<br />

REMARKS: Yamaguti (1953) described Lamellodiscus<br />

flexuosus from the gills of Synagris taeniopterus<br />

(=Nemipterus hexodon) collected at<br />

Macassar, Celebes. Young (1969) transferred this<br />

helminth to Calydiscoides Young, 1969, based on<br />

the presence of "telescoping lamellae" in the lamellodisc,<br />

and Oliver (1987) recognized Young's<br />

reassignment of L. flexuosus to Calydiscoides. In<br />

Copyright © 2011, The Helminthological Society of Washington


158 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Figures 43-52. Calydiscoides flexuosus (Yamaguti, 1953) Young, 1969. 43. Whole mount (composite,<br />

ventral; dorsal lamellodisc not shown), showing positions of hook pairs. 44, 45. Copulatory complexes. 46.<br />

Enlargement of reproductive organs (composite, ventral). 47. Dorsal bar. 48. Ventral bar. 49. Hook. 50.<br />

Dorsal anchor. 51, 52. Ventral anchors. All figures are drawn to the 25-fj.m scale, except Figures 43 and<br />

46 (200-jJiin and 50-(xm scales, respectively).<br />

addition, Oliver (1987) transferred Lamellospina<br />

indiana Karyakarte and Das, 1978, to Calydiscoides<br />

as C. indianus, considered Lamellospina<br />

Karyakarte and Das, 1978, a junior synonym of<br />

Calydiscoides, and placed C. indicus Venkatanarsaiah<br />

and Kulkarni, 1980, in synonymy with C.<br />

Copyright © 2011, The Helminthological Society of Washington<br />

52<br />

indianus. Present findings support Young's (1969)<br />

and Oliver's (1987) taxonomic proposals. Our examination<br />

of the holotype and paratypes of L.<br />

flexuosus and new voucher specimens of C. flexuosus<br />

collected from the Indian coast and Kuwait<br />

confirmed that L. flexuosus Yamaguti, 1953, and


KRITSKY ET AL.—DIPLECTANIDS FROM KUWAIT 159<br />

Table 2. <strong>Comparative</strong> measurements (in micrometers) of Calydiscoides flexuosus (Yamaguti, 1953)<br />

Young, 1969, from 3 species of Nemipterus (Nemipteridae) from the Persian Gulf and Indian Ocean.<br />

Body<br />

Length<br />

Width<br />

Haptor<br />

Length<br />

Width<br />

Lamellodisc<br />

Length<br />

Width<br />

Pharynx<br />

Width<br />

Copulatory organ<br />

Length<br />

Accessory piece<br />

Length<br />

Dorsal anchor<br />

Length<br />

Base width<br />

Ventral anchor<br />

Length<br />

Base width<br />

Bar length<br />

Dorsal<br />

Ventral<br />

Hook<br />

Length<br />

Germarinum<br />

Width<br />

Testis<br />

Length<br />

Width<br />

N. peronii Kuwait<br />

771 (<strong>67</strong>5-843; n = 8)<br />

142 (110-162; n= 11)<br />

94 (78-103; /; = 10)<br />

107 (97-114; /; = 10)<br />

49 (37-60; n = 10)<br />

41 (34-46; n = 10)<br />

42 (35-48; /;<br />

32 (29-37; n = 6)<br />

21 (17-26; n = 6)<br />

35 (32-40; n = 14)<br />

11 (10-12; n = 4)<br />

46 (43-48; n = 14)<br />

21(1 6-26; n = 12)<br />

45 (39-54; n = 14)<br />

54 (45-65; // = ID<br />

12 (11-13; n = 9)<br />

28 (24-34; n = 3)<br />

149 (123-177;<br />

63 (42-76; n<br />

= 11)<br />

n = 9)<br />

= 9)<br />

N. bipunctatus Kuwait<br />

776 (680-987; n = 9)<br />

134 (107-162: n = 9)<br />

102 (78-128; n = 9)<br />

104 (88-1 16; n = 7)<br />

49 (43-58; n = 7)<br />

37 (31-45; n - 6)<br />

38 (33-44; n = 8)<br />

36 (32-39; n = 7)<br />

19(1 5-24; n = 5)<br />

35 (30-39; n - 13)<br />

10 (9-1 1; n =<br />

46 (43-52; n = 14)<br />

22 (19-25; /; = 10)<br />

43 (37-51; n = 14)<br />

52 (47-58; n = ID<br />

11 (10-13; // = 6)<br />

29 (26-33; n = 3)<br />

149 (121-178;<br />

70 (51-89; n<br />

Calydiscoides indianus (Karyakarte and Das,<br />

1978) Oliver, 1987, and its synonyms Lamellospina<br />

Indiana Karyakarte and Das, 1978, and Calydiscoides<br />

indicus Venkatanarsaiah and Kulkarni,<br />

1980, are synonyms of C. flexuosus (Yamaguti,<br />

1953) Young, 1969.<br />

Specimens of Calydiscoides flexuosus from<br />

India, Kuwait, and the Celebes are morphologically<br />

indistinguishable. However, we did observe<br />

some differences in dimensions of the<br />

body and haptoral sclerites (Table 2). Specimens<br />

from the Celebes were somewhat smaller than<br />

those from India, while those from Kuwait were<br />

intermediate in size. These differences are not<br />

considered sufficient to separate the collections<br />

: 2)<br />

n = 9)<br />

= 9)<br />

N. japonicus India<br />

893 (781-1,044; n = 17)<br />

136 (100-187; n = 19)<br />

126 (113-142; n = 12)<br />

121 (91-149; n '=<br />

8)<br />

68 (54-76; n =<br />

42 (36-53; n =<br />

48 (23-56; n = 14)<br />

36 (29-48; n =<br />

11 (9-14; n =<br />

49 (42-56; n =<br />

21 (18-25; n =<br />

48 (38-66; n =<br />

56 (49-64; n =<br />

12-13 (/i =<br />

41 (37-45; n =<br />

13)<br />

16)<br />

27)<br />

10)<br />

30)<br />

13)<br />

18)<br />

10)<br />

13)<br />

: 4)<br />

183 (149-219; n = 9)<br />

79 (62-93; n = 9)<br />

N. hcxodon Celebes<br />

528 (411-796;<br />

74 (62-9 1 ; n<br />

88 (69-113; n = 17)<br />

93 (76-119; n = 16)<br />

43 (31-56; n = 15)<br />

31 (27-36; n = 20)<br />

23 (19-27; n = 18)<br />

27 (25-29; n = 19)<br />

9 (8-10; n =<br />

39 (35-43; n = 21)<br />

17 (15-20; n = 3)<br />

37 (34-42; n = 25)<br />

43 (39-48: n = 16)<br />

11 (10-12; n = 9)<br />

—<br />

—<br />

n = 16)<br />

= 20)<br />

: 2)<br />

into distinct species, and could result from effects<br />

of different environmental and host factors<br />

on the parasite. All previous descriptions of this<br />

species lack detail and clarity of the morphological<br />

features necessary to identify the species;<br />

our redescription provides details of the morphology<br />

of the sclerotized parts of the haptor<br />

and copulatory complex.<br />

Protolamellodiscus senilobatus sp. n.<br />

(Figs. 53-60)<br />

DESCRIPTION (measurements of specimens<br />

from A. filamentosus follow those from the type<br />

host in brackets): Lamellodiscinae. Body<br />

1,065 (720-1318; n = 8) [714 (<strong>67</strong>3-755; n =<br />

Copyright © 2011, The Helminthological Society of Washington


160 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Figures 53-60. Protolamellodiscus senilobatus sp. n. 53. Whole mount (composite, dorsal; ventral lamellodisc<br />

not shown). 54. Hook. 55. Copulatory complex. 56. Ventral anchor. 57. Ventral bar. 58. Dorsal<br />

bar. 59. Dorsal anchor. 60. Dorsal view of haptor showing dorsal lamellodisc and positions of hook pairs<br />

(ventral lamellodisc not shown). All figures are drawn to the 25-fJim scale, except Figures 53 and 60 (100fxm<br />

and 50-u.m scales, respectively).<br />

Copyright © 2011, The Helminthological Society of Washington<br />

60


2)] long, slender, fusiform; greatest width 185<br />

(120-240; n = 9) [164 (148-179; n = 2)] at<br />

level of testis. Tegument smooth. Cephalic margin<br />

narrow; 2 terminal, 2 bilateral cephalic lobes<br />

poorly developed; 3 bilateral pairs of head organs<br />

with anterior, posterior pairs associated<br />

with respective cephalic lobes; cephalic glands<br />

posterolateral to pharynx. Eyes 4; members of<br />

posterior pair slightly larger, closer together than<br />

anterior members; anterior pair frequently absent;<br />

granules irregular, variable in size; accessory<br />

granules common in cephalic region.<br />

Mouth subterminal, ventral to pharynx; pharynx<br />

73 (61-89, n = 10) [60 (51-70; n = 2)] wide,<br />

ovate or somewhat truncated posteriorly; esophagus<br />

short to nonexistent; intestinal ceca blind.<br />

Peduncle narrow, elongate. Haptor 206 (169-<br />

235; n = 8) [155 (150-161; n = 2)] wide, 111<br />

(104-117; n = 8) [84 (79-89; n = 2)] long, with<br />

3 bilateral pairs of lobes containing respective<br />

hook pairs 2, 3, 4 near apices; anterior lobes<br />

about half the length of more posterior lobes;<br />

lamellodiscs similar, each 44 (37-53; n = 10)<br />

[37 (35-39; n = 2)] long, 32 (29-38; n = 10)<br />

[30 (29-31; n = 2)] wide, with 1 complete, 8<br />

incomplete lamellae lacking medial indentation;<br />

lamellae appear to telescope somewhat in dorsoventral<br />

view. Ventral anchor 45 (38-49; n =<br />

11) [42-43 (n = 1)] long, with elongaite roots<br />

(deep root longest), evenly curved shaft, point<br />

acutely recurved not reaching level of tip of superficial<br />

root; base 14 (9-16; n = 8) wide. Dorsal<br />

anchor 41 (37-44; n = 17) [35-36 (n = 1)]<br />

long, with elongate deep root, short thickened<br />

superficial root, straight shaft, point reaching<br />

past level of tip of superficial root; base 9 (8-<br />

10; n = 13) wide. Ventral bar 41 (34-47; n =<br />

17) [36 (34-38; n = 2)] long, plate-like, with<br />

short knob-like ends; dorsal bar 40 (35-46; n<br />

= 23) [36 (34-38; n = 3)] long, with medial<br />

bend, spinous projection at proximal end.<br />

Hooks similar; each 10 (9-11; n = 28) [9-10<br />

(n = 3)] long, with protruding slightly depressed<br />

thumb, delicate point, shank; hook pair<br />

1 lying near base of ventral anchor; pairs 2, 3,<br />

4 at apices of respective haptoral lobes; pair 5<br />

posterior to ends of ventral bar; pair 6 near<br />

point of dorsal anchor; pair 7 near base of dorsal<br />

anchor. FH loop nearly shank length. Copulatory<br />

complex comprising articulated male<br />

copulatory organ, accessory piece. Male copulatory<br />

organ 45 (38-53; n = 22) [42-43 (n =<br />

1)] long, a curved heavily sclerotized tube with<br />

KRITSKY ET AL.~DIPLECTANIDS FROM KUWAIT 161<br />

subterminal recurved spine, distal loop terminating<br />

broadly; base of male copulatory organ<br />

lacking sclerotized margin. Accessory piece 28<br />

(18-34; n = 16) [32-33 (n = 1)] long, comprising<br />

flattened proximal portion, bifurcating<br />

near midlength to terminally acute elongately<br />

striated branch, spatulate branch frequently<br />

folded upon itself distally. Testis 107 (101-113;<br />

n = 2) long, 52 (48-55; n = 2) wide, ovate;<br />

vas deferens looping left intestinal cecum; seminal<br />

vesicle fusiform, simple dilation of vas deferens,<br />

lying slightly to left of body midline;<br />

prostatic reservoir saccate, lying anterior to<br />

copulatory complex. Ovary 47 (44-56; n = 5)<br />

wide, pyriform, looping right intestinal cecum,<br />

lying transversely to diagonally anterior to testis;<br />

oviduct elongate; ootype, uterus not observed;<br />

vaginal aperture sinistrodorsal, submarginal;<br />

vagina short, nonsclerotized, with proximal<br />

chamber containing apparent spermatophore,<br />

opening into medial seminal receptacle;<br />

vitellaria dense throughout trunk, except absent<br />

in regions of reproductive organs. One egg (deformed<br />

during mounting) infrequently present<br />

in uterus, with short proximal filament.<br />

TYPE HOST: King soldierbream, Argyrops<br />

spinifer (Forsskal, 1775) (Sparidae).<br />

TYPE LOCALITY: Persian Gulf off Kuwait (15<br />

January 1994).<br />

INFECTION SITE: Gills.<br />

OTHER RECORD: Soldierbream, Argyrops filamentosus<br />

(Valenciennes, 1830) (Sparidae): Persian<br />

Gulf off Kuwait (18 October 1995).<br />

SPECIMENS STUDIED: Holotype, USNPC<br />

89005; 28 paratypes from A. spinifer, USNPC<br />

89006, HWML 15021; 3 voucher specimens<br />

from A. filamentosus, USNPC 89027.<br />

ETYMOLOGY: The specific name is from Latin<br />

(sen/i — six + lobat/o = lobe) and refers to<br />

the 6 bilateral lobes of the haptor.<br />

REMARKS: Oliver (1987) recognized 3 species<br />

of Protolamellodiscus from hosts of 3 marine<br />

teleost families: Protolamellodiscus serranelli<br />

(Euzet and Oliver, 1965) Oliver, 1969, from<br />

the comber, Serranus cabrilla (Linnaeus, 1758),<br />

the brown comber, Serranus hepatus (Linnaeus,<br />

1758), and the painted comber, Serranus scriba<br />

(Linnaeus, 1758), Serranidae; Protolamellodiscus<br />

convolutus (Yamaguti, 1953) Oliver, 1987,<br />

from N. hexodon, Nemipteridae; and Protolamellodiscus<br />

raibauti Oliver and Radujkovic,<br />

1987, from the annular seabream, Diplodus annularis<br />

(Linnaeus, 1758), Sparidae. The fourth<br />

Copyright © 2011, The Helminthological Society of Washington


162 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

species, P. senilobatus sp. n., occurs on sparid<br />

hosts (Argyrops spp.)- The new species most<br />

closely resembles P. raibauti in the comparative<br />

morphology of the copulatory complex but differs<br />

from this species by possessing a subterminal<br />

spine arising from the male copulatory organ,<br />

3 bilateral pairs of haptoral lobes (lobes<br />

lacking in P. raibauti), a flattened subrectangular<br />

ventral bar (bar rod-shaped in P. raibauti),<br />

and each dorsal bar with a proximal spine (see<br />

Oliver and Radujkovic, 1987). Protolamellodiscus<br />

senilobatus differs from P. serranelli in the<br />

comparative morphology of the copulatory complex.<br />

While Yamaguti's (1953) description of<br />

P. convolutus lacks details of the sclerotized<br />

structures of the haptor and copulatory complex,<br />

P. senilobatus is distinguished from this species<br />

by possessing 3 bilateral pairs of haptoral lobes.<br />

Oliver and Radujkovic (1987) described the<br />

vagina of P. raibauti as opening sublaterally on<br />

the left side of the body. In P. senilobatus, the<br />

vaginal aperture is submarginal on the sinistrodorsal<br />

body surface, midway between the ovary<br />

and copulatory complex. In P. senilobatus, the<br />

vas deferens loops the left intestinal cecum,<br />

while Euzet and Oliver (1965) reported the vas<br />

deferens to be intercecal in P. serranelli. Oliver<br />

and Radujkovic (1987) did not observe the<br />

course of the vas deferens relative to the intestine<br />

in P. raibauti. Confirmation of these 2 characters<br />

as potential diagnostic features of Protolamellodiscus<br />

is required.<br />

Members of Protolamellodiscus Oliver, 1969,<br />

and Calydiscoides Young, 1969, are characterized,<br />

in part, by having a ventral and a dorsal<br />

lamellodisc, each with several concentric unpaired<br />

lamellae, with the most anterior lamella<br />

forming a complete circle. Calydiscoides is, in<br />

part, diagnosed by the presence of telescoping<br />

lamellae. Depending on the orientation of the<br />

lamellodisc when examined microscopically,<br />

specimens of P. senilobatus occasionally show<br />

that the deeper lamellae telescope, although not<br />

to the extent exhibited in described species of<br />

Calydiscoides. While outside the scope of the<br />

present study, it is possible that Protolamellodiscus<br />

and Calydiscoides are synonyms. Further<br />

study of all species in these genera combined<br />

with a phylogenetic analysis is necessary to clarify<br />

synonymy and/or validity of the genera.<br />

Discussion<br />

In his revision of the Diplectanidae, Oliver<br />

(1987) divided the family into 4 subfamilies<br />

Copyright © 2011, The Helminthological Society of Washington<br />

based primarily on the morphology and presence/absence<br />

of the accessory adhesive organs<br />

of the haptor. He recognized the Diplectaninae<br />

Monticelli, 1903 ("squamodiscs" composed of<br />

concentric rows of sclerotized rodlets): Lamellodiscinae<br />

Oliver, 1969 ("lamellodiscs" composed<br />

of concentric lamellae); Rhabdosynochinae<br />

Oliver, 1987 (lateral "placodiscs" unarmed);<br />

and Murraytrematoidinae Oliver, 1982<br />

(accessory adhesive organs absent).<br />

Oliver (1987) removed the then monotypic<br />

Rhamnocercinae Monaco, Wood, and Mizelle,<br />

1954, from the Diplectanidae, elevated it to familial<br />

level, and placed it in the poorly supported<br />

superfamily Heterotesioidea Euzet and DOS-<br />

SOU, 1979 (see Kritsky and Boeger, 1989), apparently<br />

because some previous descriptions of<br />

species of Rhamnocercus stated that the intestinal<br />

ceca are "apparently" united posterior to the<br />

gonads (Hargis, 1955, in R. bairdiella Hargis,<br />

1955; subsequently by Luque and lannacone<br />

[1991] in R. oliveri Luque and lannacone, 1991).<br />

However, Monaco et al. (1954) and Seamster<br />

and Monaco (1956) did not mention the intestine<br />

in the respective descriptions of R. rhamnocercus<br />

Monaco, Wood, and Mizelle, 1954, and<br />

R. stichospinus Seamster and Monaco, 1956.<br />

Luque and lannacone (1991) stated that the intestinal<br />

ceca end blindly in Rhamnocercoides<br />

menticirrhi Luque and lannacone, 1991, and<br />

Rhamnocercus stelliferi Luque and lannacone,<br />

1991. It appears that errors have been made concerning<br />

the morphology of the gut in some species<br />

of Rhamnocercinae, and the value of this<br />

character in determining familial relationships is<br />

limited. Along with Diplectaninae, Lamellodiscinae,<br />

Rhabdosynochinae, and Murraytrematoidinae,<br />

we tentatively consider the Rhamnocercinae<br />

a member of the Diplectanidae, based on<br />

general haptoral and internal morphology. However,<br />

these subfamilies all lack evolutionary support<br />

(phylogenetic analyses are lacking), and<br />

some or all may be unnatural.<br />

With the exception of Diplectanum Diesing,<br />

1858, and Lamellodiscus Johnston and Tiegs,<br />

1922, all diplectanid genera are defined by derived<br />

autapomoiphic features, suggesting that<br />

Diplectanum and Lamellodiscus are unnatural<br />

(paraphyletic) and currently serve as "catchall"<br />

groups for species lacking obvious derived characters.<br />

Kritsky and Boeger (1989) and Kritsky<br />

and Kulo (1992) discussed the probability of the


creation of paraphyletic taxa when new taxa are<br />

based primarily on autapomorphic features.<br />

Oliver (1987) considered Diplectanwn to include<br />

species having a squamodisc composed of<br />

concentric U-shaped rows of rodlets. Other diplectanine<br />

genera were diagnosed with characters<br />

thought to be lacking in Diplectanwn, such<br />

as closed circular rows of rodlets (Cycloplectanum<br />

Oliver, 1968 [=Pseudorhabdosynochus Yamaguti,<br />

1958, see Kritsky and Beverley-Burton<br />

1986]), divergent rows of rodlets (Heteroplectanum<br />

Rakotofiringa, Oliver, and Lambert,<br />

1987), lateral intestinal diverticula (Latericaecum<br />

Young, 1969), a row of elongate spines posterior<br />

to the squamodisc (Lepidotrema Johnston<br />

and Tiegs, 1922), 1 "squamodisc" (Monoplectanum<br />

Young, 1969), modified anchors (Pseudodiplectanum<br />

Tripathi, 1957), and parallel rodlets<br />

(Pseudolamellodiscus Yamaguti, 1953).<br />

Similarly, Oliver (1987) defined Lamellodiscus<br />

by species having paired (apparently incomplete)<br />

lamellae forming the lamellodiscs. The remaining<br />

genera in the Lamellodiscinae include<br />

forms with the following features absent in species<br />

of Lamellodiscus'. Calydiscoides Young,<br />

1969, with species having unpaired telescoping<br />

lamellae; Furnestinia Euzet and Audoin, 1959,<br />

with species lacking 1 "lamellodisc"; Protolamellodiscus<br />

Oliver, 1969, with species having<br />

closed or "O-shaped" lamella in the lamellodisc;<br />

and Telegamatrix Ramalingam, 1955, with<br />

a reproductive appendix containing the copulatory<br />

complex and vagina. Some of these genera<br />

might not be valid, as suggested by the apparent<br />

close relationship of Diplectanum cazauxi with<br />

Laterocaecum pearsoni and Protolamellodiscus<br />

senilobatus with species of Calydiscoides (see<br />

remarks under D. cazauxi and P. senilobatus).<br />

That Diplectanum and Lamellodiscus are paraphyletic<br />

is supported by observations on specimens<br />

in the present study. Both genera include<br />

species with varying characters, which were not<br />

considered generic features by Oliver (1987),<br />

but that could be used to determine monophyletic<br />

groups in the 2 genera. Features such as<br />

presence/absence of an accessory piece, position<br />

of the vaginal aperture, and morphology of the<br />

copulatory complex, among others, may have<br />

value in determining monophyletic groupings in<br />

these genera. In Lepidotrema, which is characterized<br />

by species possessing a posterior shield<br />

of elongate spines as its autapomorphic character,<br />

at least 1 species, L. longipenis, apparently<br />

KRITSKY ET AL.—DIPLECTANIDS FROM KUWAIT 163<br />

lacks these structures. Thus, even some of the<br />

unique characters defining some of these genera<br />

(posterior spinous shield in Lepidotrema; gut diverticula<br />

in Laterocaecum) may not be valid for<br />

defining monophyletic groups within the Diplectanidae.<br />

Acknowledgments<br />

The authors are grateful to Dr. J. R. Lichtenfels<br />

(USNPC) and Dr. J. Araki (MPM) for allowing<br />

access to type and voucher specimens in<br />

their care. Dr. N. Agarwal, Department of Zoology,<br />

University of Lucknow, Lucknow, India,<br />

contributed specimens of Calydiscoides flexuosus<br />

from the western Indian coast for use in the<br />

present study.<br />

Literature Cited<br />

Euzet, L., and G. Oliver. 1965. Lamellodiscus serranelli<br />

n. sp. (Monogenea) parasite de Teleosteens<br />

du genre Serranus. Annales de Parasitologie Humaine<br />

et Comparee 40:261-264.<br />

Hargis, W. J. 1955. Monogenetic trematodes of Gulf<br />

of Mexico fishes. Part III. The superfamily Gyrodactyloidea.<br />

Quarterly Journal of the Florida<br />

Academy of Sciences 18:33-47.<br />

Hay ward, C. J. 1996. Revision of diplectanid monogeneans<br />

(Monopisthocotylea, Diplectanidae) in<br />

sillaginid fishes, with a description of a new species<br />

of Monoplectanum. Zoologica Scripta 25:<br />

203-213.<br />

Johnston, T. H., and O. W. Tiegs. 1922. New gyrodactyloid<br />

trematodes from Australian fishes, together<br />

with a reclassification of the super-family<br />

Gyrodactyloidea. Proceedings of the Linnean Society<br />

of New South Wales 47:83-131.<br />

Karyakarte, P. P., and S. R. Das. 1978. A new<br />

monogenetic trematode, Lamellospina Indiana<br />

n. gen., n. sp. (Monopisthocotylea: Diplectanidae)<br />

from the marine fish, Nemipterus japonicus (Gunther)<br />

in India. Rivista di Parassitologia 39:19-22.<br />

Kritsky, D. C., and M. Beverley-Burton. 1986. The<br />

status of Pseudorhabdosynochm Yamaguti, 1958,<br />

and Cycloplectanum Oliver, 1968 (Monogenea:<br />

Diplectanidae). Proceedings of the Biological Society<br />

of Washington 99:17-20.<br />

, and W. A. Boeger. 1989. The phylogenetic<br />

status of the Ancyrocephalidae Bychowsky, 1937<br />

(Monogenea: Dactylogyroidea). Journal of <strong>Parasitology</strong><br />

75:207-211.<br />

, and S.-D. Kulo. 1992. Schilbetrematoides<br />

pseudodactylogyrus gen. et sp. n. (Monogenoidea,<br />

Dactylogyridae, Ancyrocephalinae) from the gills<br />

of Schilbe intermedium (Siluriformes, Schilbeidae)<br />

in Togo, Africa. Journal of the Helminthological<br />

Society of Washington 59:195-200.<br />

, V. E. Thatcher, and W. A. Boeger. 1986.<br />

Neotropical Monogenea. 8. Revision of Urocleidoides<br />

(Dactylogyridae, Ancyrocephalinae). Pro-<br />

Copyright © 2011, The Helminthological Society of Washington


164 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

ceedings of the Helminthological Society of<br />

Washington 53:1-37.<br />

Luque, J. L., and J. lannacone. 1991. Rhamnocercidae<br />

(Monogenea: Dactylogyroidea) in sciaenid<br />

fishes from <strong>Peru</strong>, with description of Rhamnocercoides<br />

menticirrhi n. gen., n. sp. and two new species<br />

of Rhamnocercus. Revista de Biologia Tropical<br />

39:193-201.<br />

Mizelle, J. D. 1936. New species of trematodes from<br />

the gills of Illinois fishes. American Midland Naturalist<br />

17:785-806.<br />

, and A. R. Klucka. 1953. Studies on monogenetic<br />

trematodes. XIV. Dactylogyridae from<br />

Wisconsin fishes. American Midland Naturalist<br />

49:720-733.<br />

, and C. E. Price. 1963. Additional haptoral<br />

hooks in the genus Dactylogyrus. Journal of <strong>Parasitology</strong><br />

49:1028-1029.<br />

Monaco, L. H., R. A. Wood, and J. D. Mizelle. 1954.<br />

Studies on monogenetic trematodes. XVI. Rhamnocercinae,<br />

a new subfamily of Dactylogyridae.<br />

American Midland Naturalist 52:129-132.<br />

Murray, F. V. 1931. Gill trematodes from some Australian<br />

fishes. <strong>Parasitology</strong> 23:492-506.<br />

Oliver, G. 1974. Nouveaux aspects du parasitisme des<br />

Diplectanidae Bychowsky, 1957 (Monogenea,<br />

Monopisthocotylea) chez les Teleosteens Perciformes<br />

des cotes de France. Comptes Rendus de<br />

1'Academic des Sciences (Paris, Serie D) 279:<br />

803-805.<br />

. 1987. Les Diplectanidae Bychowsky, 1957<br />

(Monogenea, Monopisthocotylea, Dactylogyridea)<br />

Systematique: biologie: ontogenie: ecologie: essai<br />

de phylogenese. Doctor of Science Thesis,<br />

1'Universite des Sciences et Techniques du Languedoc,<br />

Academic de Montpellier, France, 433 pp.<br />

, and I. Paperna. 1984. Diplectanidae Bychowsky,<br />

1957 (Monogenea, Monopisthocotylea),<br />

parasites de Perciformes de Mediterranee orientale,<br />

de la mer Rouge et de 1'ocean Indien. Bul-<br />

Obituary Notice<br />

MICHAEL J. PATRICK<br />

March 9, 1962-March 10, <strong>2000</strong><br />

Elected to Membership in 1989<br />

Copyright © 2011, The Helminthological Society of Washington<br />

letin du Museum National d'Historic Naturelle,<br />

Paris, 4th series, 6 (section A, No. l):49-65.<br />

, and B. Radujkovic. 1987. Protolamellodiscus<br />

raibauti n. sp., une nouvelle espece de Diplectanidae<br />

Bychowsky, 1957 (Monogenea, Monopisthocotylea)<br />

parasite de Diplodus annularis<br />

(Linnaeus, 1758) (Sparidae). Annales de Parasitologie<br />

Humaine et Comparee 62:209-213.<br />

Rakotofiringa, S., and C. Maillard. 1979. Helminthofaune<br />

des Teleostei de Madagascar: revision du<br />

genre Pseudolamellodiscus Yamaguti, 1953<br />

(Monogenea). Annales de Parasitologie (Paris) 54:<br />

507-518.<br />

Seamster, A., and L. H. Monaco. 1956. A new species<br />

of Rhamnocercinae. American Midland Naturalist<br />

55:180-183.<br />

Sey, O., and F. M. Nahhas. 1997. Digenetic trematodes<br />

of marine fishes from the Kuwaiti coast of<br />

the Arabian Gulf: Family Monorchiidae Odhner,<br />

1911. Journal of the Helminthological Society of<br />

Washington 64:1-8.<br />

Tripathi, Y. R. 1957. Studies on the parasites of Indian<br />

fishes. II. Monogenea, family: Dactylogyridae.<br />

Indian Journal of Helminthology 7:5—24.<br />

Venkatanarsaiah, J., and T. Kulkarni. 1980. New<br />

monogenetic trematode of the genus Calydiscoides<br />

Young, 1969 from the gills of Neinipterus japonicus.<br />

Proceedings of the Indian Academy of<br />

<strong>Parasitology</strong> 1:20-22.<br />

Yamaguti, S. 1934. Studies on the helminth fauna of<br />

Japan. Part 2. Trematodes of fishes, 1. Japanese<br />

Journal of Zoology 5:249-541.<br />

. 1953. Parasitic worms mainly from Celebes.<br />

Part 2. Monogenetic trematodes of fishes. Acta<br />

Medicinae Okayama 8:203-256 (with 9 plates).<br />

Young, P. C. 1969. Some monogenoideans of the<br />

family Diplectanidae Bychowsky, 1957 from Australian<br />

teleost fishes. Journal of Helminthology 43:<br />

223-254.


Comp. Parasitol.<br />

<strong>67</strong>(2), 20(K) pp. 165-168<br />

Langeronia burseyi sp. n. (Trematoda: Lecithodendriidae) from the<br />

California Treefrog, Hyla cadaverina (Anura: Hylidae), with Revision<br />

of the Genus Langeronia Caballero and Bravo-Hollis, 1949<br />

MURRAY D. DAiLEY1-3 AND STEPHEN R. GOLDBERG2<br />

1 The Marine Mammal Center, Marin Headlands, Sausalito, California, U.S.A. 94965<br />

(e-mail: daileym@tmmc.org) and<br />

2 Department of Biology, Whittier <strong>College</strong>, Whittier, California, U.S.A. 90608 (e-mail:<br />

sgoldberg @ whittier.edu)<br />

ABSTRACT: Langeronia burseyi sp. n. (Trematoda: Lecithodendriidae), a new trematode from the small intestine<br />

of Hyla cadaverina Cope, 1866, is described and illustrated. One (0.03%) of 36 adult specimens of H. cadaverina<br />

collected from Orange County, California, U.S.A., harbored 83 specimens of L. burseyi sp. n. Langeronia bursevi<br />

sp. n. is distinguished from all other species in the genus by body size, location of the cirrus, length of the ceca,<br />

placement of the vitellaria, and the shape of the excretory bladder. This is the first report of a species of<br />

Langeronia from a member of the Hylidae. An emended diagnosis and key to the genus Langeronia are presented.<br />

KEY WORDS: Digenea, Lecithodendriidae, Langeronia burseyi, new species description, taxonomy, California<br />

treefrog, Hyla cadaverina, Orange County, California, U.S.A.<br />

The taxonomic statuses of the genera Langeronia<br />

Caballero and Bravo-Hollis, 1949, and<br />

Loxogenes Stafford, 1904, have been the subject<br />

of much controversy. Caballero and Bravo-Hollis<br />

(1949) erected the genus Langeronia for a<br />

new species, Langeronia macrocirra, from the<br />

northern leopard frog, Rana pipiens Schreber,<br />

1782, in Mexico. A second species, Langeronia<br />

provitellaria, was described by Sacks (1952)<br />

from the Florida leopard frog, Rana sphenocephala<br />

Cope, 1886, in Florida, U.S.A. Yamaguti<br />

(1958) considered Langeronia synonymous with<br />

the genus Loxogenes Stafford, 1905. Brenes et<br />

al. (1959) examined specimens recovered from<br />

the cane toad, Bufo marinus (Linnaeus, 1758) in<br />

Costa Rica and disagreed with Yamaguti, concluding<br />

that Langeronia was a valid genus. Ubelaker<br />

(1965) collected trematodes from B. marinus<br />

in Nicaragua and published a redescription<br />

of L. macrocirra, concluding that L. provitellaria<br />

should be considered a synonym of that species.<br />

He also supported Yamaguti and his 1958<br />

synonymy of the 2 genera. Christian (1970)<br />

studied specimens collected from the intestines<br />

of R. pipiens in Wisconsin, Ohio, and Vermont,<br />

U.S.A., which he identified as L. provitellaria,<br />

Loxogenes sp., and a new species, Langeronia<br />

parva, respectively. Christian (1970) disagreed<br />

with Yamaguti's (1958) opinion synonymizing<br />

3 Corresponding author.<br />

165<br />

Langeronia and Loxogenes and supported Brenes<br />

et al. (1959) in validating the generic status<br />

of Langeronia. Christian (1970) did not mention<br />

the article by Ubelaker (1965) and the synonomy<br />

of L. macrocirra and L. provitellaria. However,<br />

Christian (1970) did state that in his opinion,<br />

according to the description and measurements<br />

given by Brenes et al. (1959) for "L. macrocirra,"<br />

they were actually redescribing L.<br />

provitellaria. This would tend to explain why<br />

Ubelaker (1965) synonymized the 2 species,<br />

comparing the overlap of measurements from<br />

his specimens with the measurements given by<br />

Brenes et al. (1959). Yamaguti (1958) cited Loxogenes<br />

s. str. and Langeronia as subgenera of<br />

the genus Loxogenes s. lat. Babero and Golling<br />

(1974) reported 3 species of L. provitellaria<br />

from 2 bullfrogs (Rana catesbiana Shaw, 1802)<br />

collected in Nye County, Nevada, U.S.A.<br />

Materials and Methods<br />

One of 36 California treefrogs, Hyla cadaverina<br />

Cope, 1866, examined (LACM No. 88937) from<br />

Orange County, California was infected with 83 trematodes<br />

in the large intestine. All H. cadaverina specimens<br />

had been collected between 1952 to 19<strong>67</strong> and<br />

deposited in the herpetology collection of the Natural<br />

History Museum of Los Angeles County (LACM).<br />

They were originally preserved in 10% formalin and<br />

later stored in 70% ethanol.<br />

Worms were removed from the large intestine,<br />

rinsed in 70% ethanol, stained in Delafield's hematoxylin,<br />

dehydrated in ethanol, and mounted in Canada<br />

Copyright © 2011, The Helminthological Society of Washington


166 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

balsam. Subsequent examination of the trematode<br />

specimens indicated that they represented an undescribed<br />

species of the genus Langeronia. Drawings were<br />

made with the aid of a drawing tube. Measurements<br />

are in micrometers unless otherwise indicated. The<br />

range is followed by the mean in parentheses. Type<br />

specimens were deposited in the United <strong>State</strong>s National<br />

Parasite Collection (USNPC), Beltsville, Maryland,<br />

U.S.A.<br />

Some of the cotype specimens of L. macrocirra<br />

(USNPC No. 37127), the type specimens of L. provitellaria<br />

(USNPC No. 47569), and the type and paratype<br />

specimens of L. parva (USNPC No. 70557,<br />

70558) were examined during this study.<br />

Description<br />

Results<br />

Langeronia burseyi sp. n.<br />

(Fig. 1)<br />

Based on 10 of 83 specimens: Lecithodendriidae<br />

(Liihe, 1901) Odhner, 1910; Pleurogenetinae<br />

Travassos, 1921. Body small, pyriform,<br />

0.60-0.75 mm (0.66) long, maximum width<br />

0.38-0.55 mm (0.49) at testicular level. Tegument<br />

thin, spinose. Oral sucker subterminal, 95—<br />

105 (102) long by 70-93 (81) wide. Prepharynx<br />

absent. Pharynx 60-68 (63) long by 38-45 (42)<br />

wide. Esophagus 23-28 (24) long by 12-15 (13)<br />

wide. Ceca bifurcate just anterior to midbody<br />

and extending posteriorly to anterior of testes.<br />

Acetabulum approximating size of oral sucker,<br />

75-105 (92) long by 78-98 (90) wide. Cirrus<br />

pouch 225-287 (261) long by 53-70 (59) wide,<br />

arching transversely over acetabulum, then<br />

twisting medioventrally and opening into shallow<br />

thin-walled atrium with genital pore. Testes<br />

smooth, opposite, transversely oval, in posterior<br />

third of body. Right testis 92-155 (119) long by<br />

98-163 (138) wide, left testis 110-125 (113)<br />

long by 125-188 (146) wide. Ovary round to<br />

oval, at acetabular level, anterior to right testis,<br />

56-90 (68) long by 56-100 (77) wide. Seminal<br />

receptical ovoid to spherical, 70 long by 45<br />

wide. Mehlis' gland directly postacetabular,<br />

Laurer's canal not observed. Vitellaria dorsal,<br />

follicular, extending from just posterior to pharynx<br />

to anterior half of ceca on either side of<br />

esophagus. Uterus with irregular transverse<br />

loops filling post-testicular space. Eggs smooth,<br />

elliptical, 23-28 (24) long by 12-15 (13) wide.<br />

Excretory bladder V-shaped, excretory pore terminal.<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Taxonomic summary<br />

TYPE HOST: California treefrog, Hyla cadaverina<br />

Cope, 1866, deposited in Natural History<br />

Museum of Los Angeles County as LACM<br />

88937.<br />

TYPE LOCALITY: Harding Canyon, Orange<br />

County, California, U.S.A. (33°42'N,<br />

117°38'W).<br />

COLLECTION DATE: 16 June 1965.<br />

SITE OF INFECTION: Large intestine.<br />

DEPOSITED SPECIMENS: Holotype and paratypes<br />

USNPC No. 89628<br />

ETYMOLOGY: This species is named for<br />

Charles R. Bursey, Pennsylvania <strong>State</strong> University,<br />

Shenango, Pennsylvania, U.S.A., in recognition<br />

of his many contributions to the parasitology<br />

of amphibians and reptiles.<br />

Langeronia Caballero and Bravo-Hollis,<br />

1949<br />

EMENDED DIAGNOSIS: Lecithodendriidae,<br />

Pleurogenetinae. Body spatulate to pyriform,<br />

spined. Oral sucker well-developed, terminal or<br />

subterminal. Prepharynx present or absent; pharynx<br />

well-developed. Esophagus present; ceca<br />

wide, extending to midbody. Acetabulum equatorial.<br />

Testes symmetrical, postacetabular, and<br />

intercecal. Cirrus pouch elongate, twisted, extending<br />

transversely intercecally in space between<br />

intestinal bifurcation and acetabulum.<br />

Genital pore preacetabular, ventral to or on internal<br />

border of left cecum. Ovary dextral to acetabulum,<br />

pretesticular. Uterine coils lateral,<br />

postequatorial; eggs smooth, operculate. Vitellaria<br />

follicular, in shoulder area, on either side<br />

of esophagus, not confluent. Excretory bladder<br />

Y- or V-shaped. Intestinal parasites of amphibians.<br />

TYPE SPECIES: Langeronia macrocirra Caballero<br />

and Bravo-Hollis, 1949, from R. pipiens<br />

in Mexico.<br />

OTHER SPECIES: In addition to L. macrocirra,<br />

the genus currently contains 2 other species,<br />

Langeronia provitellaria Sacks, 1952, and L.<br />

parva Christian, 1970. Langeronia burseyi sp. n.<br />

differs from all members of the genus in its<br />

small size (smallest in the genus), placement of<br />

cirrus at the acetabular level, cecal length, and<br />

V-shaped bladder. It most closely resembles L.<br />

provitellaria in the position of the vitellaria,<br />

with both beginning at the pharyngeal level.<br />

However, in L. burseyi the vitellaria end just


DAILEY AND GOLDBERG—LANGERONIA BURSEYI SP. N. FROM TREEFROGS 1<strong>67</strong><br />

Figure 1. Langeronia burseyi sp. n. from Hyla cadaverina. Entire worm, ventral view.<br />

posterior to the cecal bifurcation, while in L. (size, position of vitellaria and pharynx, length<br />

provitellaria, they extend to the anterior of the of ceca) distinguish L. macrocirra and L. procirrus.<br />

The new species resembles L. macrocirra vitellaria as separate species,<br />

and L. parva in that the ovary and testes are not<br />

deeply lobed as in L. provitellaria. The deeply Key to the Species of Langeronia<br />

lobed ovary and testes along with other features la. Body length more than 1.30 mm 2<br />

Copyright © 2011, The Helminthological Society of Washington


168 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

lb. Body length less than 1.30 mm 3<br />

2a. Ovary and testes deeply lobed .. L. provitellaria<br />

2b. Ovary and testes not deeply lobed<br />

L. macrocirra<br />

3a. Prepharynx present L. parva<br />

3b. Prepharynx absent L. burseyi<br />

Remarks and Discussion<br />

During this study, the specimens treated by<br />

Ubelaker (1965) and Christian (1970) were not<br />

available. However, after examination of all other<br />

known available specimens, we agree with<br />

Christian (1970) that Langeronia is a valid genus<br />

and that L. macrocirra and L. provitellaria<br />

are valid species. The differences between Loxogenes<br />

and Langeronia are as follows: In Loxogenes,<br />

the vitellaria are confluent, while in<br />

Langeronia they are not. In Loxogenes, the testes<br />

are on the same level as the acetabulum, and<br />

the ovary is always preacetabular; while in Langeronia,<br />

the testes are postacetabular, and the<br />

ovary is at the same level or just postacetabular.<br />

In Loxogenes, the cirrus pouch is extracecal and<br />

club-shaped, and the genital pore is extracecal,<br />

anterior and dorsal. In Langeronia, the cirrus<br />

pouch is elongate, not club-shaped, twisted intercecally,<br />

and preacetabular, while the genital<br />

pore is lateral and ventral to the left cecum. In<br />

Loxogenes, the intestinal ceca do not extend to<br />

the acetabulum, while in Langeronia, they always<br />

extend past the acetabulum. In Loxogenes,<br />

the uterine coils are arranged in an anterior-posterior<br />

configuration, while in Langeronia, the<br />

uterine coils are lateral loops confined to the<br />

posterior half of the body.<br />

Acknowledgments<br />

The authors thank Robert L. Bezy, Natural<br />

History Museum of Los Angeles County, for<br />

Obituary Notice<br />

ALAN F. BIRD<br />

February 11, 1928-December 13, 1999<br />

Elected to Honorary Membership in 1997<br />

Copyright © 2011, The Helminthological Society of Washington<br />

permission to examine Hyla cadaverina; J.<br />

Ralph Lichtenfels, United <strong>State</strong>s National Parasite<br />

Collection, for the loan of type material; and<br />

Lynn Hertel, University of New Mexico, for her<br />

help with illustrations.<br />

Literature Cited<br />

Babero, B. B., and K. Coiling. 1974. Some helminth<br />

parasites of Nevada bullfrogs, Rana catesbiana<br />

Shaw. Revista de Biologia Tropical, Universidad<br />

de Costa Rica 21:207-220.<br />

Brenes, R. R., G. Arroyo-Sancho, and E. Delgado-<br />

Flores. 1959. Helmintos de la Republica de Costa<br />

Rica XI. Sobre la validez del genero Langeronia<br />

Caballero y Bravo, 1949 (Trematoda: Lecithodendriidae)<br />

y hallazgo de Ochetosoma miladelarocai<br />

Caballero y Vogelsang, 1947. Revista de<br />

Biologfa Tropical, Universidad de Costa Rica 7:<br />

81-87.<br />

Caballero, E., and M. Bravo-Hollis. 1949. Description<br />

d'un nouveau genre de Pleurogeninae (Trematoda:<br />

Lecithodendriidae) de grenouilles du Mexique<br />

(1) Langeronia macrocirra n. g., n. sp. Annales<br />

de Parasitologie Humaine et Comparee 24:<br />

193-199.<br />

Christian, F. A. 1970. Langeronia pai~va sp. n. (Trematoda:<br />

Lecithodendriidae) with a revision of the<br />

genus Langeronia Caballero and Bravo-Hollis,<br />

1949. Journal of .<strong>Parasitology</strong> 56:321-324.<br />

Sacks, M. 1952. Langeronia provitellaria (Lecithodendriidae),<br />

a new species of trematode from Rana<br />

pipiens sphenocephala. Transactions of the American<br />

Microscopical Society 71:2<strong>67</strong>—269.<br />

Ubelaker, J. E. 1965. The taxonomic status of Langeronia<br />

Caballero and Bravo-Hollis, 1949 with<br />

the synonymy of Loxogenes provitellaria Sacks,<br />

1952 with Loxogenes macrocirra Caballero and<br />

Bravo-Hollis, 1949. Transactions of the Kansas<br />

Academy of Science 68:187-190.<br />

Yamaguti, S. 1958. Systema Helminthum. Vol. 1. The<br />

Digenetic Trematodes of Vertebrates, Parts I and<br />

II. Interscience Publishers, Inc., New York.


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 169-180<br />

Oxyuroids of Palearctic Testudinidae: New Definition of the Genus<br />

Thaparia Ortlepp, 1933 (Nematoda: Pharyngodonidae), Redescription<br />

of Thaparia thapari ihapari, and Descriptions of Two New Species<br />

SALAH BouAMER1-3 AND SERGE MoRAND2<br />

1 Groupe Pluridisciplinaire des Sciences, Universite de Perpignan, Av. de Villeneuve, 66860 Perpignan, France<br />

(e-mail: bouamer@univ-perp.fr) and<br />

2 Centre de Biologie et d'Ecologie Tropicale et Mediteraneenne, Laboratoire de Biologic Animale (UMR 5555<br />

CNRS), Universite de Perpignan, Av. de Villeneuve, 66860 Perpignan, France (e-mail: morand@univ-perp.fr)<br />

ABSTRACT: The generic diagnosis of Thaparia Ortlepp, 1933, is emended based on the study and redescription<br />

of Thaparia thapari thapari (Dubinina, 1949) from the cecum of Testudo graeca Linnaeus, 1758, collected in<br />

Settat, Morocco. In addition, 2 new species, Thaparia carlosfeliui sp. n. and Thaparia bourgati sp. n. from the<br />

cecum of Testudo hermanni Gmelin, 1789, collected in Catalonia, Spain, are described. Scanning electron<br />

microscopy studies revealed substantial differences in the structure of the mouth and the caudal end, which<br />

enabled us to differentiate the 2 new species from the others and from each other.<br />

KEY WORDS: Thaparia thapari thapari, Thaparia carlosfeliui sp. n., Thaparia bourgati sp. n., Nematoda,<br />

Pharyngodonidae, Testudo graeca, spur-thighed tortoise, Testudo hermanni, Hermann's tortoise, Morocco, Spain.<br />

The genus Thaparia was erected by Ortlepp<br />

(1933) for Thaparia macrospiculum Ortlepp,<br />

1933, a parasite of the tent tortoise, Psammobates<br />

tentorius (Bell, 1828). Ortlepp (1933) gave<br />

the following diagnosis: Medium-sized worms<br />

possessing 3 lips and a relatively short esophagus<br />

consisting of an anterior muscular portion,<br />

a middle glandular portion, and a posterior bulb;<br />

excretory pore post-bulbar; lateral alae absent.<br />

Vulva approximated to anus; vagina very long;<br />

2 uteri and 2 ovaries. Caudal extremity of male<br />

cut ventrally and continued backward to form a<br />

short truncated and alate tail. Four pairs of caudal<br />

papillae, 3 pairs circumcloacal and 1 pair<br />

toward tip of tail. Single spicule very long and<br />

stout, extending to or even anterior of the esophageal<br />

bulb. Type species T. macrospiculum from<br />

P. tentorius.<br />

Walton (1942) described a second species,<br />

Thaparia contortospicula, a parasite of the Galapagos<br />

tortoise, Geochelone nigra (Quoy and<br />

Gaimard, 1824) from the Galapagos Islands.<br />

Fitzsimmons (1961) described Thaparia capensis<br />

from the South African bowsprit tortoise,<br />

Chersina angulata (Schweigger, 1812), in South<br />

Africa.<br />

Fetter (1966) described Thaparia domerguei,<br />

from the common spider tortoise, Pyxis arachnoides<br />

Bell, 1827, and the radiated tortoise,<br />

Geochelone radiata (Shaw, 1802), from Mada-<br />

Corresponding author.<br />

169<br />

gascar. She also transferred the species Tachygonetria<br />

thapari Dubinina, 1949, described from<br />

the Central Asian tortoise, Testudo horsfieldii<br />

Gray, 1844, in Afghanistan and from other Palearctic<br />

tortoises, the spur-thighed tortoise Testudo<br />

graeca Linnaeus, 1758, and Hermann's tortoise,<br />

Testudo hermanni Gmelin, 1789, to the<br />

genus Thaparia. Fetter (1966) also modified the<br />

diagnosis of the genus, which is now: Pharyngodoninae—mouth<br />

with 3 lips; short esophagus<br />

divided into 2 equal parts; 4 pairs of caudal papillae:<br />

3 pairs at the level of the cloaca and 1<br />

near tail extremity. Caudal alae present or absent<br />

in males. Type species: T. macrospiculum Ortlepp,<br />

1933.<br />

Petter and Douglass (1976) described 2 additional<br />

species, T. rnacrocephala and T. microcephala,<br />

from the Bolson tortoise, Gopherus flavomarginatus<br />

Legler, 1959, in Mexico.<br />

Baker (1987) cited the 7 species above plus 3<br />

subspecies of T. thapari, following Petter (1966)<br />

(Thaparia thapari thapari (Dubinina, 1949),<br />

Thaparia thapari australis (Petter, 1966), and<br />

Johnson (1973a) (Thaparia thapari lysavyi<br />

Johnson, 1973, from Testudo hermanni from Albania).<br />

The 7 species of Thaparia are distributed<br />

in 5 biogeographical regions.<br />

In this study, we describe 2 new species of<br />

Thaparia from a testudinid species and emend<br />

the diagnosis of the genus.<br />

Materials and Methods<br />

A first collection of nematode parasites from 18<br />

specimens of Testudo graeca from Settat, Morocco<br />

Copyright © 2011, The Helminthological Society of Washington


170 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

(deposited at the Institut Agronomique et Veterinaire<br />

Hassan II, Rabat, Morocco), was made by one of us<br />

(S.B.). The second collection, from a single specimen<br />

of T. hennanni from Catalonia, Spain, was made by<br />

C. Feliu, Barcelona, Spain, and deposited at the Barcelona<br />

Zoo, Spain. Nematodes were preserved in 70%<br />

ethanol before being cleared with lactophenol for<br />

study. Figures were made with the aid of a drawing<br />

tube. Nematodes were dehydrated by passage through<br />

progressive ethanol concentrations to absolute ethanol<br />

and critical-point-dried (M scope 500, Hitachi, Japan).<br />

The scanning electron microscope used was a Hitachi<br />

S 520, Hitachi, Japan at 20 kV. Measurements given<br />

are for the holotype male and the allotype female.<br />

Measurements in parentheses are the ranges of paratype<br />

males and females. All measurements are in micrometers.<br />

Results<br />

Thaparia thapari thapari (Dubinina, 1949)<br />

(Figs. 1-10)<br />

Redescription<br />

GENERAL: The material examined consisted<br />

of 6 males and 15 females. Body medium-sized,<br />

stout. Mouth triangular, with 3 transparent lips.<br />

Buccal cavity with denticles. Cephalic sense organs<br />

consisting of inner circle of 6 nerve endings,<br />

papillae not pedunculate (Figs. 2, 3, 6, 7),<br />

the outer circle not observed, and amphids present.<br />

Esophagus divided into 2 portions: anterior<br />

muscular part, and comparatively longer posterior<br />

glandular part terminating in valvular bulb;<br />

17 chitinoid pieces surrounding anterior end of<br />

esophagus (Figs. 1, 3). Excretory pore postesophageal.<br />

MALE: Length 2,754-3,169; maximum<br />

thickness 191-229. In worm measuring 2,773,<br />

esophagus 388: corpus 180 and isthmus plus<br />

bulb 208. Nerve ring and excretory pore 161<br />

and 889, respectively, from anterior end. Posterior<br />

extremity truncated. Tail <strong>67</strong> long. Spicule<br />

needle-shaped, 100 long. Gubernaculum Vshaped.<br />

Three pairs of caudal papillae: 2 circumcloacal<br />

(1 pair preanal and 1 pair postanal) and<br />

1 pair at tail end. Preanal membrane, present<br />

with 6 lobes (Figs. 4, 5, 8, 9, 10), and caudal<br />

alae absent.<br />

FEMALE: Length 4,282-4,716; maximum<br />

thickness 356-378. In a worm measuring 4,600,<br />

esophagus 615: corpus 240 and isthmus plus<br />

bulb 375. Nerve ring, excretory pore, and vulva<br />

at 180, 1,282, and 2,264, respectively, from anterior<br />

end. Tail 270 long.<br />

Taxonomic summary<br />

HOST: Spur-thighed tortoise, Testudo graeca<br />

Linnaeus, 1758.<br />

SITE IN HOST: Cecum.<br />

TYPE LOCALITY/COLLECTION DATE: Settat,<br />

Morocco, 32°30'45"N, 7°45'30"W, 22 July 1999,<br />

by S.B.<br />

SPECIMENS DEPOSITED: Museum National<br />

d'Histoire Naturelle, Paris, France. Number 825<br />

HE<br />

Remarks<br />

Thapar (1925) described the female of this<br />

species from Testudo graeca, as Oxyuris sp. Dubinina<br />

(1949) studied males, the structures of<br />

which made it possible to include the species in<br />

Tachygonetria Wedl, 1862. Fetter (1966) redescribed<br />

and transferred this species to the genus<br />

Thaparia and divided it in 2 subspecies: T. thapari<br />

thapari and T. thapari australis.<br />

The emended diagnosis characterizes T. thapari<br />

thapari with 17 chitinoid pieces, whereas<br />

Petter (1961, 1966) cited 6 chitinoid pieces. Cephalic<br />

sense organs consist of an inner circle of<br />

6 nerve endings (papillae not pedunculate), 4<br />

submedian and 2 lateral close to amphids; outer<br />

papillae were not observed.<br />

Description<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Thaparia carlosfeliui sp. n.<br />

(Figs. 11-23)<br />

GENERAL: The material examined consisted<br />

of 20 males and 40 females. Nematoda, Oxyuroidea,<br />

Pharyngodonidae, Thaparia. Robust<br />

worms of small size. Mouth surrounded by 3<br />

lips. Esophagus in 2 parts, with elongated isthmus.<br />

Amphids prominent. Differs from the diagnosis<br />

of the genus in the number of caudal<br />

papillae.<br />

MALE: (holotype and 3 paratypes): Mouth<br />

surrounded by 3 V-shaped cut lips (Figs. 13, 21).<br />

Six oral papillae, arranged in 3 pairs (Fig. 13).<br />

Buccal cavity without denticles. Esophagus<br />

lobes visible. No chitinoid pieces visible. Tail<br />

without alae. Structure of caudal region complex<br />

(Figs. 15, 16, 22, 23). One pair of preanal papillae,<br />

1 pair of large postanal elongated papillae.<br />

Posterior lip of anus with central nipple.<br />

Surrounding ventral membrane present lateral<br />

and anterior to anus, and second preanal membrane<br />

situated posterior to first membrane. Anterior<br />

lip of anus with 2 lobes, extremity of spic-


BOUAMER AND MORAND—OXYUROID GENUS THAPARIA 171<br />

Figures 1-5. Thaparia thapari thapari, male. 1. Anterior end. 2. Cephalic end, en face view. 3. Cephalic<br />

end, deeper en face view. 4. Posterior end, lateral view. 5. Posterior end, ventral view. All scale lines = 50 |un.<br />

Copyright © 2011, The Helminthological Society of Washington


172 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Figures 6-9. Thaparia thapari thapari, scanning electron micrographs: 6. Cephalic end of female. 7.<br />

Cephalic end of male. 8. Caudal end of male, ventral view (a = preanal papillae, b = postanal papillae,<br />

c = caudal papillae). 9. Caudal end of male, lateral view (a = preanal papillae, b = postanal papillae, c<br />

= caudal papillae, e = preanal membrane). Scale lines: Fig. 6 = 17.6 (xm; Fig. 7 = 20 urn; Fig. 8 = 20<br />

fjim; Fig. 9 = 20 (xm.<br />

Copyright © 2011, The Helminthological Society of Washington


Figure 10. Thaparia thaparia thaparia, scanning<br />

electron micrograph: Cloacal view of male (d = anterior<br />

lip, e = preanal membrane, f = posterior lip, g<br />

= ventral lobe of preanal membrane, h = subventral<br />

lobe of preanal membrane, i = spicule). Scale line =<br />

5 (xm.<br />

ule visible between them. U-shaped gubernaculum.<br />

Third pair of papillae lateral at end of tail.<br />

Length 1,797 (1,324-1,943) with maximal width<br />

180 (150-255). Nerve ring 112 (70-116) from<br />

anterior end. Excretory pore 586 (435-640)<br />

from anterior end. Esophagus 416 (351-421)<br />

long, corpus 165 (139-1<strong>67</strong>) long, isthmus plus<br />

bulb 251 (212-254) long. Tail 45 (45-49) long.<br />

Spicule 123 (109-123) long.<br />

FEMALE (allotype and 3 paratypes): Mouth<br />

surrounded by 3 V-shaped cut lips (Figs. 19, 20).<br />

Six oral papillae arranged in 3 pairs as in male<br />

(Fig. 19). Buccal cavity without denticles. No<br />

chitinoid pieces visible. Esophagus lobes visible.<br />

Length 3,471 (2,8<strong>67</strong>-3,848), maximum width<br />

426 (426-482). Nerve ring corpus 114 (114-<br />

188) from anterior end. Excretory pore 814<br />

(575-814) from anterior end. Esophagus: 615<br />

(482-634) long with corpus 244 (192-252) long<br />

and isthmus plus bulb 371 (290-382). Vulva<br />

1,622 (1,528-2,056) from anterior end. Tail 244<br />

(232-283) long. Eggs asymmetrical, measuring<br />

89 X 147 (69 X 126-89 X 147).<br />

BOUAMER AND MORAND—OXYUROID GENUS THAPARIA 173<br />

Taxonomic summary<br />

TYPE HOST: Hermann's tortoise, Testudo hermanni<br />

Gmelin, 1789.<br />

SITE IN HOST: Cecum.<br />

TYPE LOCALITY/COLLECTION DATE: South Catalonia,<br />

Spain, 41°23'14"N, 2°11'17"E, 17 December<br />

1993 by Dr. Carlos Feliu.<br />

SPECIMENS DEPOSITED: Museum National<br />

d'Histoire Naturelle, Paris, France. Number 826<br />

HE<br />

ETYMOLOGY: The species is named in honor<br />

of Professor Carlos Feliu (University of Barcelona,<br />

Spain).<br />

Remarks<br />

Thaparia carlosfeliui sp. n. differs from T.<br />

macrospiculum in the size of the spicule and<br />

from T. domerguei in the absence of caudal alae.<br />

Thaparia capensis differs from the new species<br />

in the presence of caudal alae and the length of<br />

the body.<br />

Thaparia contortospicula resembles T. carlosfeliui<br />

in the size of the spicule but differs in<br />

the presence of caudal alae and the length of the<br />

body. Thaparia macrocephala and T. microcephala<br />

differ in the presence of caudal alae, the<br />

number of lips, and the length of the body.<br />

Thaparia carlosfeliui differs from T. thapari<br />

(Dubinina, 1949) in the absence of teeth, the arrangement<br />

of labial papillae, and the absence of<br />

6 chitinoid pieces around the mouth. Thaparia<br />

carlosfeliui differs from T. thapari australis in<br />

the arrangement of labial papillae, the lack of a<br />

tip at the extremity of the male tail, and the absence<br />

of chitinoid pieces.<br />

Thaparia bourgati sp. n.<br />

(Figs. 24-31)<br />

Description<br />

GENERAL: The material examined consisted<br />

of 10 males and 1 female. Nematoda, Oxyuroidea,<br />

Pharyngodonidae, Thaparia. Robust<br />

worms of small size. Mouth surrounded by 3<br />

lips. Esophagus in 2 parts, with an elongated<br />

isthmus. Amphids prominent. Differs from the<br />

diagnosis of the genus in the number of caudal<br />

papillae.<br />

MALE (holotype and 3 paratypes): Labial papillae<br />

conspicuous, arranged as in T. carlosfeliui<br />

sp. n. (Figs. 20, 28). Three projections visible<br />

inside mouth, forming diaphragm under lip.<br />

Buccal cavity without denticles. No chitinoid<br />

Copyright © 2011, The Helminthological Society of Washington


174 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

15<br />

Copyright © 2011, The Helminthological Society of Washington


BOUAMER AND MORAND—OXYUROID GENUS THAPARIA 175<br />

Figures 17-19. Thaparia carlosfeliui sp. n. female allotype. 17. Entire specimen, lateral view. 18. Anterior<br />

end. 19. Cephalic end, en face view. Scale lines: Fig. 17 = 400 (Jim; Figs. 18, 19 = 50 (xm.<br />

pieces visible. Tail without alae. Three pairs of<br />

caudal papillae. Ventral membrane surrounding<br />

anus anteriorly, showing on each side 2 small<br />

submedian lobes and 2 large lateral lobes (Figs.<br />

26, 27, 29—31). Anterior lip of anus composed<br />

of 2 submedian lobes congruent at distal extremity.<br />

Posterior lip of anus with central nipple. Ushaped<br />

gubernaculum. Length 3,122 (2,204-<br />

3,207), maximum width 221 (143-255). Nerve<br />

ring 98 (68-104) from anterior end. Excretory<br />

pore 756 (705-851) from anterior end. Esophagus<br />

439 (251-416) long, corpus 169 (97-169)<br />

long, isthmus plus bulb 270 (154-270) long. Tail<br />

53 (45-53) long. Spicule 162 (112-225) long.<br />

FEMALE (allotype): A single female has been<br />

found, which closely resembles females of the<br />

other species of the genus. Nerve ring not visible.<br />

Length 2,226, maximum width 166. Excre-<br />

Figures 11-16. Thaparia carlosfeliui sp. n. male holotype. 11. Entire specimen, lateral view. 12. Anterior<br />

end. 13. Cephalic end, en face view. 14. Cephalic end, deeper en face view. 15. Posterior end, lateral<br />

view. 16. Posterior end, ventral view. Scale lines: Fig. 11 = 100 jim; Figs. 12-16 = 50 jim.<br />

Copyright © 2011, The Helminthological Society of Washington


176 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

zz Z3<br />

Figures 20-23. Thaparia carlosfeliui sp. n., scanning electron micrographs. 20. Cephalic end of female.<br />

21. Cephalic end of male. 22. Caudal end of male, lateral view (a = preanal papillae, b = postanal papillae,<br />

c = caudal papillae, d = anterior lip, e = preanal membrane). 23. Caudal end of male, ventral view (a =<br />

preanal papillae, b = postanal papillae, c = caudal papillae, d = anterior lip, e = preanal membrane,<br />

f = posterior lip). Scale lines: Fig. 20 = 30 urn; Figs. 21—23 = 25 jjim.<br />

tory pore 549. Esophagus 502, corpus 225 long,<br />

isthmus plus bulb 277 long. Vulva 804 from anterior<br />

end. Tail 131. Eggs asymmetrical, measuring<br />

49 X 90.<br />

Taxonomic summary<br />

TYPE HOST: Hermann's tortoise, Testudo hermanni<br />

Gmelin, 1789.<br />

Copyright © 2011, The Helminthological Society of Washington<br />

INFECTION SITE: Cecum.<br />

TYPE LOCALITY/COLLECTION DATE: South Catalonia,<br />

Spain, 41°23'14°N, 2°11'17"E, 17 December<br />

1993, by Dr. Carlos Feliu.<br />

SPECIMENS DEPOSITED: Museum National<br />

d'Histoire Naturelle, Paris, France. Number 827<br />

HE<br />

ETYMOLOGY: The species is named in honor<br />

c


BOUAMER AND MORAND—OXYUROID GENUS THAPARIA 177<br />

Figures 24-27. Thaparia bourgati sp. n. male holotype. 24. Cephalic end, en face view. 25. Cephalic<br />

end, deeper en face view. 26. Caudal end, lateral view. 27. Caudal end, ventral view. All scale lines = 50<br />

|xm.<br />

Copyright © 2011, The Helminthological Society of Washington


178 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Figures 28—31. Thaparia bourgati sp. n. male, scanning electron micrograph. 28. Cephalic end. 29.<br />

Caudal end, lateral view (a = preanal papillae, b = postanal papillae, c = caudal papillae). 30. Caudal<br />

end, ventral view (a = preanal papillae, b = postanal papillae, c = caudal papillae, d = anterior lip, e =<br />

preanal membrane). 31. Cloacal view (d = anterior lip, e = preanal membrane, f = posterior lip). Scale<br />

lines: Fig. 28 = 10 fjim; Fig. 29 = 17.6 jjim; Fig. 30 = 15 u.m; Fig. 31 = 7.5 fjim.<br />

Copyright © 2011, The Helminthological Society of Washington


of Professor Robert Bourgat (University of Perpignan,<br />

France).<br />

Remarks<br />

Thaparia bourgati sp. n. differs from all other<br />

species of the genus in the same characters as<br />

T. carlosfeliui. Thaparia bourgati differs from<br />

T. carlosfeliui in the shape of the preanal membrane<br />

in the male—9 lobes in T. bourgati and 4<br />

lobes in T. carlosfeliui—and in the size of the<br />

eggs (smaller in T. bourgati).<br />

Discussion<br />

Thaparia bourgati sp. n. and T. carlosfeliui<br />

sp. n. differ from all other species of the genus<br />

Thaparia, except T. thapari, in the lack of caudal<br />

alae. Both species differ from the subspecies<br />

T. thapari australis in the shape and the disposition<br />

of labial papillae, the lack of apical chitinoid<br />

pieces, the lack of a tip at the end of the<br />

male tail, and the presence of a longer spicule<br />

compared with the length of the tail. They differ<br />

from the subspecies T. thapari thapari in the<br />

disposition of labial papillae, the lack of esophageal<br />

teeth and the lack of apical chitinoid pieces,<br />

the shape of the preanal membrane, and the<br />

shape of the gubernaculum in the male, and<br />

from T. thapari rysavyi in the arrangement of<br />

labial papillae and in the shape of the adanal<br />

BOUAMER AND MORAND—OXYUROID GENUS THAPARIA 179<br />

tudinidae. Medium-sized, lateral alae present or<br />

absent. Mouth with 3 or 6 slightly bilobed lips.<br />

Esophagus rather short, divided into 2 parts of<br />

about equal length: an anterior muscular corpus<br />

and a posterior glandular isthmus terminating in<br />

a valvulated bulb. Excretory pore bulbar or postbulbar.<br />

MALE: Tail truncated, spiked, or simple.<br />

Caudal alae present or absent. Spicule simple or<br />

contorted, may be very long. Gubernaculum U-,<br />

V-, or Y-shaped. Caudal papillae in 3 pairs: 2<br />

circumcloacal, 1 pair at or near tail end.<br />

FEMALE: Tail tapering to sharp point. Vulva<br />

postequatorial, sometimes very close to anus.<br />

Vagina long; ovijector present. Eggs thinshelled,<br />

relatively few.<br />

TYPE SPECIES: Thaparia macrospiculum Ortlepp,<br />

1933, in Psammobates tentorius; South<br />

Africa.<br />

Key to the Species and Subspecies of the<br />

Genus Thaparia<br />

This key follows Johnson (1973b) and includes<br />

T. capensis Fitzsimmons, 1961, T. macrocephala<br />

Petter and Douglass (1976), T. microcephala<br />

Petter and Douglass (1976), and T.<br />

carlosfeliui sp. n. and T. bourgati sp. n., both<br />

described herein.<br />

membrane in the male. Finally, T. carlosfeliui 1. Caudal alae in male present 2<br />

sp. n. resembles T. bourgati sp. n. in the arrange- Caucal alae in male absent ..... 7<br />

ment of the labial papillae, but it is distinguished 2' Iai! in ma!e spiked " I<br />

^ r Tail in male truncated 5<br />

by the shape of the preanal membrane in the 3 Spicule contorted; vulva away from anus<br />

male and by the size of the eggs. T. contortospicula Walton, 1942<br />

The genus Thaparia shows a wide geograph- Spicule simple; vulva far away from anus ..... 4<br />

ical distribution, with 3 Palearctic species (T. 4- Len§th of sPicule 1/2 of body<br />

..... _, , . . „, , T. macrocephala Fetter and Douglass (1976)<br />

carlosfelim sp. n., T. bourgati sp. n., and T. tha- Length of spicule 1/3 of body<br />

pan), 2 Nearctic species, 2 South African spe- T. micmcephala Fetter and Douglass (1976)<br />

cies, and 1 species from the Galapagos Islands. 5. Spicule simple; vulva far (more than 1,450 [Jim)<br />

The question remains open concerning the pres- from anus T. capensis Fitzsimmons, 1961<br />

ence of this genus in other members of the Pa- Splcule simPle; vulva near (less than 20° ^m) f<br />

^ anus 6<br />

learctic tortoises (T. horsfieldii, T. graeca, T. 6 Spicule less than 1 mm in length<br />

hermanni, the Egyptian tortoise, Testudo klein- T. domerguei Fetter, 1966<br />

manni Lortet, 1883, and the marginated tortoise, Spicule more than 2.5 mm in length<br />

Testudo marginata Schoepff, 1792). T macrospiculum Ortlepp, 1933<br />

7. Buccal cavity with 6 teeth<br />

Emended diagnosis of the genus Thaparia , , . T: ll*apari th«pari Dubinina (1949)<br />

to l Buccal cavity without teeth 8<br />

The use of a scanning electron microscope al- g. Spicule less than 90 jxm; tail more than 90 jxm<br />

lowed verification that the previously described in length T. thapari australis Fetter, 1966<br />

adanal papillae are simple lobes. The lack of ter- Spicule more than 90 jun; tail less than 90 ^.m<br />

. , . c j u * - , u in length 9<br />

mmal nerves is confirmed by optical observa- 9 preana, ^embrane absent<br />

tion. The new diagnosis of the genus is: T. thapari rysavyi Johnson, 1973<br />

Pharyngodonidae: Intestinal parasites of Tes- Preanal membrane present 10<br />

Copyright © 2011, The Helminthological Society of Washington


180 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

10. Preanal membrane with 4 lobes<br />

T. carlosfeliui sp. n.<br />

Preanal membrane with 9 lobes<br />

T. bourgati sp. n.<br />

Acknowledgment<br />

We thank Dr. Annie Fetter for helpful comments<br />

on an earlier version of the manuscript.<br />

Literature Cited<br />

Baker, M. R. 1987. Synopsis of the Nematoda parasitic<br />

in amphibians and reptiles. Memorial University<br />

of Newfoundland, Occasional Papers in<br />

Biology 11:1-325.<br />

Dubinina, M. H. 1949. (Ecological studies on the parasite<br />

fauna of the Testudo horsfieldii Gray from<br />

Tadjikistan). Parazitologicheskii Sbornik Zoologicheskogo<br />

Instituta AN SSSR 11:61-97. (In Russian.)<br />

Fitzsimmons, W. M. 1961. Thaparia capensis n. sp.,<br />

an oxyuroid parasite of Testudo angidata. British<br />

Journal of Herpetology 3:7-12.<br />

Johnson, S. 1973a. Some oxyurid nematodes of the<br />

genera Mehdiella and Thaparia from the tortoise<br />

Testudo hermani. Folia Parasitologica 20:141—<br />

148.<br />

Obituary Notice<br />

MARION M. FARR<br />

1903-<strong>2000</strong><br />

. 1973b. The first record of the nematode Thaparia<br />

thapari thapari (Dubinina, 1949) in Afghanistan,<br />

with remarks on the genus Thaparia<br />

Ortlepp, 1933. Folia Parasitologica 20:178.<br />

Ortlepp, R. J. 1933. On some South African reptilian<br />

oxyurids. Onderstepoort Journal of Veterinary<br />

Science and Animal Industry 1:9-114.<br />

Fetter, A. J. 1961. Redescription et analyse critique<br />

de quelques especes d'Oxyures de la tortue grecque<br />

(Testudo graeca L.). Diversite des structures cephaliques.<br />

Annales de Parasitologie Humaine et<br />

Comparee 10:648-<strong>67</strong>1.<br />

. 1966. Equilibre des especes dans les populations<br />

de Nematodes parasites du colon des Tortues<br />

terrestres. Memoires du Museum National<br />

d'Histoire Naturelle, Paris, Nouvelle Serie, Serie<br />

A, Zoologie 39:1-252.<br />

, and J. F Douglass. 1976. Etude des populations<br />

d'Oxyures du colon des Gopherus (Testudinidae).<br />

Bulletin du Museum National d'Histoire<br />

Naturelle, Paris, Serie 3, No. 389, Zoologie 271:<br />

731-768.<br />

Thapar, G. S. 1925. Studies on the oxyurid parasites<br />

of the reptiles. Journal of Helmintology 3:83-150.<br />

Walton, A. C. 1942. Some oxyurids from a Galapagos<br />

tortoise. Proceedings of the Helminthological Society<br />

of Washington 9:1-17.<br />

Elected to Membership, 1938<br />

Executive Committee Member at Large, 1943-1945<br />

21st Recording Secretary, 1946<br />

Vice President, 1946<br />

34th President, 1951<br />

Assistant Secretary-Treasurer, 1964<br />

Society Representative to the Washington Academy of<br />

Sciences, 1964-1965<br />

Elected to Life Membership, 1979<br />

Published in the Proceedings from 1939-1963 on<br />

Eimeria and Histomonas<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 181-189<br />

Parasites of Farm-Raised Trout in Michigan, U.S.A.<br />

PATRICK M. MUZZALL<br />

Department of Zoology, Natural Science Building, Michigan <strong>State</strong> University, East Lansing, Michigan 48824,<br />

U.S.A. (e-mail: rnuzzall@pilot.msu.edu)<br />

ABSTRACT: A total of 635 trout (366 rainbow trout, Oncorhynchus mykiss Richardson, 1836; 166 brook trout,<br />

Salvelinus fontinalis Mitchill, 1814; 103 brown trout, Salmo trutta Linnaeus, 1758; Salmonidae) collected in<br />

March-July 1996, 1997, and 1998 from 12 trout farms in Michigan, U.S.A., was examined for parasites. Twelve<br />

parasite species (1 Acanthocephala, Acanthocephalus dims (Van Cleave, 1931) Van Cleave and Townsend, 1936;<br />

1 Monogenea, Gyrodactylus sp.; 2 Cestoda, Eubothrium salvelini (Schrank, 1790), Proteocephalus sp.; 1 Nematoda,<br />

Truttaedacnitis sp.; 1 Copepoda, Salmincola edwardsii (Olsson, 1869); 1 Myxozoa, Myxobolus cerebralis<br />

(Hofer, 1903); 4 Ciliophora, Capriniana sp. [ = Trichophrya sp.], Chilodonella sp., Ichthyophthirus multifiliis<br />

(Fouquet, 1876), Trichodina sp.; and 1 Mastigophora, Ichthyobodo sp. [ = Costia sp.]) were found. Rainbow trout<br />

were infected with A. dims, Truttaedacnitis sp., E. salvelini, Prutcocephalus sp., Gyrodactylus sp., M. cerebralis,<br />

Ichthyobodo sp., Capriniana sp., Chilodonella sp., /. multifiliis, and Trichodina sp. Brook trout were infected<br />

with A. dirus, S. edwardsii, E. salvelini, M. cerebralis, and Trichodina sp. Acanthocephalus dims was the only<br />

parasite infecting brown trout. Eubothrium salvelini, A. dirus, and Trichodina sp. infected trout from 9, 8, and<br />

8 farms, respectively. Acanthocephalus dirus in all trout species and S. edwardsii on brook trout had the highest<br />

prevalences, mean intensities, and mean abundances.<br />

KEY WORDS: trout, rainbow trout, Oncorhynchus mykiss, brook trout, Salvelinus fontinalis, brown trout, Salmo<br />

trutta, Salmonidae, helminths, protozoans, parasites, aquaculture, Michigan, U.S.A.<br />

Newman and Kevern (1994) reported that<br />

more than half of the fish growers in the state<br />

of Michigan, U.S.A., raise rainbow, brook, and<br />

brown trout. These growers produce trout for<br />

sale: 1) to individuals or groups for stocking, 2)<br />

to retail stores or restaurants, and 3) through<br />

their own fee-fishing ponds. In 1996, predators<br />

and diseases were the leading causes of death<br />

for trout in culture conditions in Michigan, accounting<br />

for 53% and 13% of all fish lost, respectively<br />

(Anonymous, 1997). Except for a few<br />

reports in local newspapers of diseases of trout<br />

in the Michigan Department of Natural Resources<br />

hatcheries and the studies by Sawyer et al.<br />

(1974) and Yoder (1972), little has been published<br />

on the parasites of trout raised in culture<br />

in Michigan. The present study reports on the<br />

parasites infecting rainbow, brook, and brown<br />

trout from 12 privately owned farms in Michigan.<br />

The emphasis of this study was on the<br />

metazoan parasites of trout, but observations and<br />

comments are also made on protozoans. Furthermore,<br />

information is presented on the life<br />

cycles of some of the parasites, their pathogenicity,<br />

and factors influencing their occurrence.<br />

Materials and Methods<br />

Trout were collected by dip net or seine from ponds<br />

or raceways (hereinafter referred to as ponds) in<br />

March-July 1996, 1997, and 1998 from 12 farms in<br />

181<br />

Michigan. These trout farms are in an area of the lower<br />

peninsula between 42.0° and 45.5°N and 84.0° and<br />

86.5°W. Specific information on the locations of the<br />

trout farms, however, cannot be provided because of<br />

conditions of confidentiality imposed by the growers.<br />

Fish were either brought to the laboratory alive and<br />

necropsied within 48 hr of collection or were put on<br />

ice at the farm, brought to the laboratory, frozen, and<br />

examined later. Total length (mm) and sex were recorded<br />

during necropsy. The fins, external surface,<br />

buccal cavity, gills, brain, eyes, gonads, swim bladder,<br />

gastrointestinal tract, liver, spleen, and muscles (left or<br />

right side of each fish) were examined from all fish.<br />

In fish collected in 1997 and 1998, the skull bones and<br />

cartilage and 2 or more gill arches (without the filaments)<br />

were macerated separately into a slurry and examined<br />

with a compound microscope at 20 X. During<br />

the entire study, rainbow trout were examined from 23<br />

ponds, brook trout from 13 ponds, and brown trout<br />

from 7 ponds. These totals include some of the same<br />

ponds sampled on different dates and in different<br />

years. Parasite prevalence is defined as the percentage<br />

of fish infected, mean intensity as the mean number of<br />

metazoan parasites in infected fish, and mean abundance<br />

as the mean number of metazoan parasites in<br />

examined fish. Population numbers of each metazoan<br />

species at 1 facility were estimated as (prevalence) X<br />

(mean abundance) X (estimated number of trout) at<br />

the facility when the fish were sampled. It should be<br />

emphasized that the results of examination of frozen<br />

trout may not accurately reflect the occurrence and (or)<br />

numbers of protozoans and monogeneans. Therefore,<br />

mean numbers were not calculated for these parasite<br />

groups. Protozoan taxonomy follows that of Lom and<br />

Dykova (1992). Voucher specimens have been deposited<br />

in the United <strong>State</strong>s National Parasite Collection<br />

Copyright © 2011, The Helminthological Society of Washington


182 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Table 1. Numbers and mean lengths of Oncorhynchus<br />

mykiss, Salvelinus fontinalis, and Salmo trutta<br />

examined in 1996, 1997, and 1998.<br />

Species*<br />

OM, 1996<br />

OM, 1997<br />

OM, 1998<br />

SF, 1996<br />

SF, 1997<br />

SF, 1998<br />

ST, 1996<br />

ST, 1997<br />

No.<br />

examined<br />

184<br />

50<br />

132<br />

112<br />

27<br />

27<br />

88<br />

15<br />

Mean length ± SD<br />

(range, 95% confidence<br />

intervals)<br />

162 ± 71 (62-338, 152-172)<br />

177 ± 59 (93-279, 161-195)<br />

215 ± 64 (94-409, 204-226)<br />

154 ± 72 (56-378, 141-168)<br />

222 ± 45 (153-328, 204-239)<br />

200 ± 26 (160-260, 190-210)<br />

146 ± 69 (43-283, 131-161)<br />

208 ± 25 (147-244, 194-222)<br />

* OM = Oncorhynchus mykiss; SF = Salvelinus fontalis', ST<br />

= Salmo trutta.<br />

(USNPC), Beltsville, Maryland, U.S.A., with the following<br />

accession numbers: Eubothrium salvelini<br />

(89483), Acanthocephalus dims (89484), Salmincola<br />

edwardsii (89485).<br />

Results<br />

Totals of 366 rainbow trout, 166 brook trout,<br />

and 103 brown trout were examined for parasites<br />

from 12 Michigan farms in March—July<br />

1996, 1997, and 1998. All farms did not raise<br />

all species; thus, unequal numbers of each species<br />

were examined. The mean lengths of the<br />

trout species examined each year are in Table 1.<br />

Rainbow trout in 1998 were significantly larger<br />

than those in 1996 and 1997 (analysis of variance,<br />

F = 24.4, P < 0.0001). Brook trout in<br />

1996 were significantly smaller that those examined<br />

in 1997 and 1998 (analysis of variance,<br />

F = 15.6, P < 0.0001). Brown trout in 1997<br />

were significantly larger than those examined in<br />

1996 (Student's f-test, t = -6.34, P < 0.0001).<br />

Twelve parasite species were found in trout in<br />

this study (Table 2). Eleven parasite species infected<br />

rainbow trout, 5 species infected brook<br />

trout, and only 1 species infected brown trout.<br />

The prevalences, mean intensities, and mean<br />

abundances of the parasites found in trout in<br />

ponds varied dramatically. Of the metazoan parasite<br />

species found, Eubothrium salvelini<br />

(Schrank, 1790), Acanthocephalus dims (Van<br />

Cleave, 1931) Van Cleave and Townsend, 1936,<br />

and Salmincola edwardsii (Olsson, 1869) were<br />

gravid. The sites (in parentheses) where the parasites<br />

were found in trout were: E. salvelini (pyloric<br />

ceca, small intestine); Proteocephalus sp.<br />

(intestine); Gyrodactylus sp. (gills), Tmttaedac-<br />

Copyright © 2011, The Helminthological Society of Washington<br />

nitis sp. (small intestine); A. dims (intestine); S.<br />

edwardsii (primarily gills, inner operculum, base<br />

of fins); Myxobolus cerebral is Hofer, 1903 (head<br />

bones and cartilage, gill arches); Ichthyobodo<br />

sp., Capriniana sp., Chilodonella sp., Ichthyophthirus<br />

multifiliis (Fouquet, 1876), and Trichodina<br />

sp. (gills and [or] head area).<br />

Acanthocephalus dims was the most common<br />

parasite species occurring in the gastrointestinal<br />

tract of each trout species. When S. edwardsii<br />

occurred, it commonly infested brook trout. Although<br />

several small immature S. edwardsii<br />

were seen on many brook trout, mean intensities<br />

and mean abundances of 5. edwardsii reflect<br />

only gravid females counted. Trichodina sp. was<br />

the most common external protozoan found on<br />

rainbow and brook trout. Myxobolus cerebralis<br />

infected 13 of 22 rainbow trout at 1 farm, 2 of<br />

15 rainbow trout at another farm, and 1 of 20<br />

brook trout at a third facility. Capriniana sp.,<br />

Chilodonella sp., /. multifiliis, and Ichthyobodo<br />

sp. each infected fish in only 1 pond.<br />

All farms had trout that were infected with at<br />

least 1 parasite species. Eubothrium salvelini, A.<br />

dims, and Trichodina sp. infected trout from 9<br />

(75%), 8 (<strong>67</strong>%), and 8 (<strong>67</strong>%) farms, respectively,<br />

of the 12 farms from which trout were examined.<br />

Acanthocephalus dims and Trichodina<br />

sp. each infected rainbow trout from 52% of the<br />

23 ponds examined (Table 3). Eubothrium salvelini,<br />

A. dims, and 5. edwardsii each infected<br />

brook trout from 31% of the 13 ponds. Acanthocephalus<br />

dims infected brown trout from 2<br />

of 7 ponds.<br />

Trout were examined for parasites from 1<br />

farm in March and July 1997. Prevalences, mean<br />

abundances, and estimated numbers of helminths<br />

varied dramatically between these<br />

months (Table 4). In March, 4 parasite species<br />

infected trout, and A. dims and 5. edwardsii<br />

were common. Acanthocephalus dims had the<br />

highest mean abundances and estimated number<br />

of helminths. In July, only A. dims infrequently<br />

infected trout.<br />

There were no significant differences in the<br />

prevalence (chi-square analysis, P > 0.05) and<br />

intensity (Mann-Whitney test, P > 0.05) of E.<br />

salvelini, A. dims, and S. edwardsii between female<br />

and male trout of each species. At the<br />

farms where these 3 parasite species were common,<br />

examined trout did not vary enough in<br />

length to determine if these parasite infections<br />

had a significant relationship with length.


Discussion<br />

Twelve parasite species (2 Cestoda, 1 Monogenea,<br />

1 Nematoda, 1 Acanthocephala, 1 Copepoda,<br />

1 Myxozoa, 1 Mastigophora, 4 Ciliophora)<br />

were found in 635 trout examined from 12<br />

farms in the present study. Parasites of these<br />

trout and their infection values varied between<br />

ponds and years, a variability characteristic of<br />

wild trout populations as well. Most if not all of<br />

these parasite species have been found in wild<br />

trout from Michigan environments (Muzzall,<br />

1984, 1986; Hernandez and Muzzall, 1998). Digenetic<br />

trematodes, however, found in wild trout<br />

by Muzzall (1984, 1986), did not infect trout<br />

from culture ponds.<br />

Of the parasitic species found in the present<br />

study, E. salvelini, A. dims, S. edwardsii, and<br />

M. cerebralis had prevalences of 50% or more<br />

in at least 1 pond and deserve further discussion.<br />

Eubothrium salvelini was a common parasite of<br />

rainbow and brook trout. It utilizes copepods as<br />

intermediate hosts and commonly infects wild<br />

salmonids in inland waters (Hernandez and<br />

Muzzall, 1998) as well as in the Great Lakes<br />

(Muzzall, 1993, 1995a, b). Muzzall (1984)<br />

found immature Eubothrium sp. in brook trout<br />

from a Michigan creek. Boyce (1969) reported<br />

that E. salvelini reduced the growth, swimming<br />

performance, and survival of salmon. Smith and<br />

Margolis (1970) suggested that this cestode<br />

caused indirect damage to young salmonids.<br />

Hendee in 1980 believed it reduced the growth<br />

of brook trout in the state of New Hampshire,<br />

U.S.A. (in Hoffman, 1999).<br />

Acanthocephalus dirus had the highest prevalences,<br />

mean intensities, and mean abundances<br />

of all parasites found. It is widespread in Michigan<br />

trout farms and is common in some natural<br />

environments of Michigan (Muzzall, 1984).<br />

Muzzall (1984) also reported that the isopod,<br />

Caecidotea intermedius Forbes, 1876, was the<br />

intermediate host for this parasite in the Rogue<br />

River. In the present study, some individuals of<br />

all 3 trout species infected with 100 worms or<br />

more from 3 farms appeared emaciated, and the<br />

head appeared large for the size of the fish. Bullock<br />

(1963) demonstrated that the most pronounced<br />

effects of A. dirus (=A. jacksoni) in<br />

rainbow and brook trout in a New Hampshire<br />

hatchery were damage to the intestinal epithelium<br />

and proliferation of connective tissue, leading<br />

to malnutrition and emaciation. Furthermore,<br />

MUZZALL—PARASITES OF TROUT 183<br />

he stated (p. 33) that "this worm seriously impairs<br />

the health of the fish." Allison (1954) discussed<br />

the advancements in prevention and<br />

treatment of parasitic diseases of fish and listed<br />

13 parasitic genera that warranted discussion.<br />

However, A. dirus was not listed, and acanthocephalans<br />

in general receive little attention in<br />

hatchery manuals about fish diseases. It is not<br />

known if A. dirus was common when Allison<br />

wrote his article or if it has become increasingly<br />

common in Michigan.<br />

The presence of S. edwardsii on brook trout<br />

and its absence from rainbow and brown trout<br />

were not unexpected, because it parasitizes only<br />

the former species (Kabata, 1969). Some brook<br />

trout infested with S. edwardsii had 1 or both<br />

opercula folded underneath itself, and the distal<br />

portions of many gill filaments showed hyperplasia<br />

and clubbing. These characteristics also<br />

occurred on uninfected trout, suggesting previous<br />

infestation by this parasite. This copepod<br />

has a direct life cycle and is a common parasite<br />

of brook trout in Michigan (Allison and Latta,<br />

1969; Muzzall, 1984, 1986). The mean intensities<br />

of S. edwardsii are higher than those found<br />

on trout in Michigan lotic environments but are<br />

comparable to the high mean intensities found<br />

on brook trout in Michigan lakes by Allison and<br />

Latta (1969). Many studies on Salmincola spp.<br />

suggest that they debilitate their hosts but may<br />

not be direct causes of trout mortality. Allison<br />

and Latta (1969) found no relationship between<br />

S. edwardsii and brook trout mortality in Michigan<br />

lakes.<br />

Owners of 2 Michigan farms told me that they<br />

had seen a parasitic copepod on the gills of rainbow<br />

trout in their ponds. However, in this study,<br />

none was found infesting rainbow trout. A copepod<br />

that infests rainbow trout is Salmincola<br />

californiensis (Dana, 1853), which is native to<br />

streams in the Pacific Northwest, U.S.A., and<br />

Canada (Kabata, 1969). Hoffman (1984) reported<br />

on its eastward movement in North America,<br />

being transferred on live fish and with shipments<br />

of trout eggs. Sutherland and Wittrock (1985)<br />

believed that 5. californiensis entered central<br />

Iowa through the importation of infested rainbow<br />

trout from a Missouri farm. They found a<br />

mean intensity of 4.6 adults and suggested that<br />

this copepod may be responsible for host mortality<br />

if fish are sufficiently stressed. Perhaps this<br />

species has made its way to Michigan but is infrequent<br />

on rainbow trout in this state.<br />

Copyright © 2011, The Helminthological Society of Washington


184 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Table 2. Prevalences, mean intensities, mean abundances of parasites found in Oncorhynchus mykiss,<br />

Salvelinus fontinalis, and Salmo trutta from farms in 1996, 1997, and 1998.<br />

Parasite<br />

Cestoda<br />

Eiibothrium salvelini<br />

Proteocephalus sp.<br />

Monogenea<br />

Gyrodactylus sp.<br />

Nematoda<br />

Truttaedacnitis sp.<br />

Acanthocephala<br />

Acanthocephalus dims<br />

Copepoda<br />

Salmincola edwardsii<br />

Myxozoa<br />

Myxobolus ccrebralis<br />

Ciliophora<br />

Capriniana sp.<br />

Chilodonella sp.<br />

Ichthyophthirus multifiliis<br />

Farm<br />

number,<br />

year<br />

4, 96<br />

4, 96<br />

5, 96<br />

5, 97<br />

6, 98<br />

8, 98<br />

9, 98<br />

11, 98<br />

3, 96<br />

12, 96<br />

1, 98<br />

12, 98<br />

4, 96<br />

1, 97<br />

11, 98<br />

11, 98<br />

5, 96<br />

2,<br />

4,<br />

4,<br />

4,<br />

5,<br />

5,<br />

5,<br />

5,<br />

7,<br />

8,<br />

10,<br />

11,<br />

5,<br />

5,<br />

5,<br />

8,<br />

5,<br />

5,<br />

5,<br />

96<br />

96<br />

96<br />

96<br />

96<br />

97<br />

98<br />

98<br />

98<br />

98<br />

98<br />

98<br />

96<br />

96<br />

97<br />

98<br />

96<br />

96<br />

97<br />

1, 96<br />

5, 96<br />

1, 97<br />

5, 97<br />

5, 97<br />

11, 98<br />

5, 97<br />

8, 98<br />

2, 96<br />

3, 96<br />

Trout<br />

species*<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

SF<br />

SF<br />

SF<br />

SF<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

SF<br />

SF<br />

SF<br />

SF<br />

ST<br />

ST<br />

ST<br />

SF<br />

SF<br />

SF<br />

SF<br />

OM<br />

OM<br />

SF<br />

OM<br />

OM<br />

OM<br />

No. examined<br />

25<br />

13<br />

15<br />

15<br />

16<br />

20<br />

18<br />

22<br />

14<br />

4<br />

19<br />

2<br />

25<br />

5<br />

22<br />

22<br />

15<br />

12<br />

13<br />

13<br />

25<br />

15<br />

15<br />

5<br />

16<br />

20<br />

20<br />

20<br />

22<br />

20<br />

16<br />

15<br />

6<br />

20<br />

15<br />

15<br />

10<br />

20<br />

12<br />

15<br />

15<br />

22<br />

20<br />

20<br />

12<br />

20<br />

Prevalence<br />

No. infected<br />

(%)<br />

1 (4)<br />

1 (8)<br />

3 (20)<br />

4(27)<br />

1 (6)<br />

14 (70)<br />

13 (72)<br />

5 (23)<br />

1 (7)<br />

1 (25)<br />

1 (5)<br />

1 (50)<br />

2(8)<br />

2(40)<br />

4(18)<br />

1 (5)<br />

3 (20)<br />

4(33)<br />

1 (8)<br />

1 (8)<br />

10 (40)<br />

15 (100)<br />

15 (100)<br />

1 (20)<br />

9(56)<br />

7(35)<br />

16 (80)<br />

1 (5)<br />

5(23)<br />

20 (100)<br />

3(19)<br />

15 (100)<br />

6 (100)<br />

20 (100)<br />

3 (20)<br />

15 (100)<br />

10 (100)<br />

20 (100)<br />

7(58)<br />

14 (93)<br />

2 (13)<br />

13 (59)<br />

1 (5)<br />

8 (40)<br />

3 (25)<br />

8 (40)<br />

Mean intensity<br />

±SD (max)<br />

1.7 ±<br />

2.0 ±<br />

2.4 ±<br />

20.7 ±<br />

2.6 ±<br />

1.5 ±<br />

93.5 ±<br />

2.3 ±<br />

44.7 ±<br />

56.5 ±<br />

3.2 ±<br />

17.9 ±<br />

98.1 ±<br />

5.6 ±<br />

42.7 ±<br />

2.0 ±<br />

46.3 ±<br />

11.7 ±<br />

36.0 ±<br />

3.3 ±<br />

75.7 ±<br />

1.0<br />

1.0<br />

1.2 (3)<br />

2.0 (5)<br />

1.0<br />

1.5 (5)<br />

44.9 (163)<br />

1.8 (5)<br />

2.0<br />

1.0<br />

1.0<br />

1.0<br />

1.0<br />

1.0<br />

0.6 (2)<br />

—<br />

1.0<br />

107.5 (209)<br />

1.0<br />

2.0<br />

1.6(5)<br />

43.0 (127)<br />

110.5 (432)<br />

22.0<br />

3.3 (11)<br />

39.4 (107)<br />

137.8 (490)<br />

1.0<br />

9.2 (22)<br />

48.6 (172)<br />

1.4 (4)<br />

43.8 (116)<br />

6.9 (20)<br />

24.3 (87)<br />

4.0 (8)<br />

80.6 (298)<br />

39.5 ± 20.9 (71)<br />

3.6 ± 3.1 (11)<br />

6.6 ± 6.4 (20)<br />

45.4 ± 31.2 (88)<br />

—<br />

—<br />

—<br />

Copyright © 2011, The Helminthological Society of Washington<br />

—<br />

— -<br />

Mean abundance<br />

±SD<br />

0.04 ± 0.20<br />

0.08 ± 0.27<br />

0.33 ± 0.82<br />

0.53 ± 1.30<br />

0.06 ± 0.25<br />

1,65 ± 1.66<br />

14.94 ± 6.63<br />

0.59 ± 1.37<br />

0.14 ± 0.54<br />

0.25 ± 0.50<br />

0.05 ± 0.23<br />

0.50 ± 0.71<br />

0.08 ± 0.28<br />

0.40 ± 0.55<br />

0.27 ± 0.63<br />

—<br />

0.20 ± 0.41<br />

31.20 ± 72.6<br />

0.08 ± 0.28<br />

0.15 ± 0.56<br />

0.92 ± 1.49<br />

44.70 ± 43.00<br />

56.50 ± 110.50<br />

4.40 ± 9.80<br />

1.81 ± 2.93<br />

6.25 ± 23.87<br />

78.40 ± 128.90<br />

0.05 ± 0.22<br />

1.27 ± 4.<strong>67</strong><br />

42.70 ± 48.60<br />

0.38 ± 1.03<br />

46.3 ± 43.8<br />

11.7 ± 6.9<br />

36.0 ± 24.3<br />

0.<strong>67</strong> ± 2.05<br />

75.7 ± 80.6<br />

39.5 ± 20.9<br />

3.6 ± 3.1<br />

3.8 ± 5.8<br />

42.4 ± 32.2<br />

—<br />

—<br />

—<br />

—<br />

—<br />


Table 2. Continued.<br />

Parasite<br />

Trichodina sp.<br />

Mastigophora<br />

Ichthyobodo sp.<br />

Farm<br />

number,<br />

year<br />

1, 96<br />

1, 96<br />

2, 96<br />

4, 96<br />

4, 96<br />

1, 97<br />

1, 97<br />

1, 97<br />

7, 98<br />

9, 98<br />

11, 98<br />

12, 98<br />

1, 96<br />

1, 96<br />

1, 97<br />

1, 98<br />

8, 98<br />

12, 98<br />

3, 96<br />

Trout<br />

species*<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

OM<br />

SF<br />

SF<br />

SF<br />

SF<br />

SF<br />

SF<br />

OM<br />

No. examined<br />

10<br />

24<br />

12<br />

13<br />

25<br />

10<br />

20<br />

5<br />

20<br />

18<br />

22<br />

11<br />

20<br />

10<br />

12<br />

19<br />

6<br />

2<br />

20<br />

MUZZALL—PARASITES OF TROUT 185<br />

Prevalence<br />

No. infected Mean intensity<br />

(%) ±SD (max)<br />

10 (100) —<br />

14 (58)<br />

3 (25)<br />

2(15)<br />

7(28)<br />

3 (30)<br />

2(10)<br />

1 (20)<br />

4 (20)<br />

10 (56)<br />

4(18)<br />

2(18)<br />

5 (25)<br />

10 (100)<br />

5 (42)<br />

5 (26)<br />

2(33)<br />

1 (50)<br />

2 (10) —<br />

OM = Oncorhynchus mykiss; SF = Salvelinus fontinalis; ST = Salmo trutta.<br />

Myxobolus cerebralis has been present in<br />

Michigan waters since at least 1968, when it was<br />

discovered in 3 commercial trout hatcheries.<br />

Yoder (1972) discussed the spread of M. cerebralis<br />

into native trout populations in the Tobacco<br />

River, Michigan, from 1 of these hatcheries.<br />

The protozoan spread down the first 6 mi of water,<br />

and factors involved in this spread were the<br />

high incidence of disease at the hatchery, abundance<br />

of susceptible trout, and trout movement.<br />

In 1998 and 1999, M. cerebralis was reported<br />

from at least 6 privately owned trout farms in<br />

Michigan. It has been suggested that it was endemic<br />

in 1 or more facilities and transferred to<br />

other facilities with infected fish or by piscivorous<br />

birds that ate infected fish. In the present<br />

study, M. cerebralis-infected trout were detected<br />

in 3 farms. Infected trout, however, did not<br />

have clinical symptoms. Furthermore, Sutherland<br />

(1999) reported that M. cerebralis has been<br />

found in fish from the Au Sable and Manistee<br />

rivers in lower Michigan.<br />

In the present study, parasites and their numbers<br />

infecting trout in a pond may dramatically<br />

change the next time the fish are sampled and<br />

examined. One reason for this is that trout are<br />

moved into and out of facilities during the year.<br />

An example of this was evident at 1 farm, when,<br />

Mean abundance<br />

±SD<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

in March 1997, 4 parasite species infected trout<br />

and 2 were common (Table 4). In July, only 1<br />

species infrequently infected trout after the infected<br />

trout were moved out and uninfected ones<br />

were moved in. Also, informing the owner of<br />

the facility on the parasites found can affect infection<br />

levels from 1 sampling date to the next.<br />

After an owner was informed that brook trout<br />

were infested with S. edwardsii, he told me that<br />

"a treatment had been done to the pond." Approximately<br />

2 months later, the prevalence and<br />

mean intensity of S. edwardsii on trout from the<br />

same pond were dramatically reduced. Another<br />

suggestion for these differences is that parasite<br />

species may exhibit a seasonal cycle in their occurrence.<br />

Hare and Frantsi (1974) found 12 parasite<br />

species, 10 parasite species, and 1 parasite species<br />

infecting, respectively, Atlantic salmon,<br />

Salmo solar Linnaeus, 1758; brook trout; and<br />

rainbow trout, in 13 Canadian hatcheries in the<br />

Maritime provinces. Hexamita salmonis (Moore,<br />

1923) Wenyon, 1926; Trichophyra piscium<br />

Buetschli, 1889; Diplostomum spathaceum (Rudolphi,<br />

1819) Braun, 1893; Acanthocephalus later<br />

alls (Leidy, 1851); and S. edwardsii were<br />

considered to be serious fish pathogens, based<br />

on the work of other authors. Buchmann and<br />

Copyright © 2011, The Helminthological Society of Washington<br />


186 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Table 3. Parasites of Oncorhynchus mykiss examined<br />

from 23 ponds, Salvelinus fontinalis from 13<br />

ponds, and Salmo trutta from 6 ponds in 1996,1997,<br />

and 1998.<br />

Cestoda<br />

No. (%) of<br />

ponds where<br />

Trout parasites<br />

Parasite species* occurred<br />

Eubothrium salvelini<br />

Proteocephalus sp.<br />

Monogenea<br />

Gyrodactylus sp.<br />

Nematoda<br />

Truttaedacnitis sp.<br />

Acanthocephala<br />

Acanthocephalus dims<br />

Copepoda<br />

Salmincola edwardsii<br />

Myxozoa<br />

Myxobolus cerebralis<br />

Ciliophora<br />

Capriniana sp.<br />

Chilodonella sp.<br />

Ichthyophthirus multifiliis<br />

Trichodina sp.<br />

Mastigophora<br />

Ichthyobodo sp.<br />

OM<br />

SF<br />

OM<br />

OM<br />

OM<br />

OM<br />

SF<br />

ST<br />

SF<br />

OMf<br />

OMt<br />

SFt<br />

OM<br />

OM<br />

OM<br />

OM<br />

SF<br />

OM<br />

9(39)<br />

4(31)<br />

3 (13)<br />

1 (4)<br />

1 (4)<br />

12 (52)<br />

4(31)<br />

2(33)<br />

4(31)<br />

2(17)<br />

1 (8)<br />

1 (25)<br />

1 (4)<br />

1 (4)<br />

1 (4)<br />

12 (52)<br />

6 (46)<br />

1 (4)<br />

* OM = Oncorhynchus mykiss; SF = Salvelinus fontinalis;<br />

ST = Salmo trutta.<br />

t OM examined from 12 ponds and SF from 4 ponds in 1997<br />

and 1998.<br />

Bresciani (1997) listed investigations performed<br />

on the parasites of farmed salmonids, and reported<br />

22 parasite species (12 protozoans and 10<br />

metazoans) infecting 805 pond-reared rainbow<br />

trout from 5 freshwater farms in Denmark.<br />

Based on these and the present studies, there is<br />

a relationship between the number of trout examined<br />

and number of parasite species found.<br />

As the number of fish examined increases, so<br />

does the number of parasite species found.<br />

Hnath (1993) suggested that a sample size of 60<br />

individuals should be examined from a population<br />

of 2,000 fish or more in a pond in order to<br />

detect a pathogen. The numbers of parasite species<br />

found in rainbow and brook trout in the present<br />

study are low compared with the numbers<br />

found by Hare and Frantsi (1974) and by Buchman<br />

and Bresciani (1997). More rainbow and<br />

brook trout were examined in those studies than<br />

in the present one.<br />

The total numbers of parasite species found<br />

in rainbow, brook, and brown trout in this study<br />

are low compared with the numbers for each<br />

species listed by Hoffman (1999) in North<br />

America. This may be explained by the artificial<br />

conditions in trout farms, which harbor very few<br />

potential intermediate invertebrate hosts. In most<br />

ponds, snails, which serve as intermediate hosts<br />

for digenetic trematodes, were never collected.<br />

In contrast, protozoans with direct life cycles are<br />

easily introduced and spread between fish. In<br />

conversations with several farmers, it was apparent<br />

that they used several "antiparasitic"<br />

drugs to treat against ectoparasitic infections<br />

when they were aware that their fish exhibited<br />

signs of infection. This treatment regime also<br />

explains the paucity of parasites. The current<br />

and increasing use of well water and spring wa-<br />

Table 4. Prevalence (P), mean abundance (MA), and estimated number (EN) of parasites from 1 farm<br />

in March and July 1997.<br />

Parasite<br />

Acanthocephalus dims<br />

Eubothrium salvelini<br />

Truttaedacnitis sp.<br />

Salmincola edwardsii<br />

Trout<br />

species<br />

(»)*<br />

SF (20)<br />

ST (20)<br />

OM (15)<br />

OM (15)<br />

OM (15)<br />

SF (20)<br />

P<br />

100<br />

100<br />

100<br />

20<br />

7<br />

100<br />

March<br />

42.7<br />

36.0<br />

44.7<br />

0.33<br />

0.20<br />

3.60<br />

MA ± SD<br />

(Max.)<br />

± 49 (172)<br />

± 24 (87)<br />

± 43 (127)<br />

± 0.82 (3)<br />

± 0.41<br />

± 3 (11)<br />

EN<br />

106,750<br />

90,000<br />

312,900<br />

462<br />

98<br />

9,000<br />

Trout<br />

species<br />

(/>)*<br />

SF(16)<br />

ST (15)<br />

OM (15)<br />

OM (15)<br />

OM (15)<br />

SF (16)<br />

SF = Salvelinus fontinalis; ST = Salmo trutta; OM = Oncorhynchus mykiss. (no. examined)<br />

Copyright © 2011, The Helminthological Society of Washington<br />

P<br />

19<br />

20<br />

0<br />

0<br />

0<br />

0<br />

July<br />

MA ± SD<br />

(Max.) EN<br />

0.38 ± 1.03 (4) 14<br />

0.<strong>67</strong> ± 2.05 (8) 27<br />

— —<br />

— —<br />

— —<br />


ter and fiberglass or concrete ponds and tanks<br />

will also reduce the incidence of parasites.<br />

The effects of natural bodies of water and the<br />

fish in them serving as a source of parasites in<br />

culture should be addressed. This involves facilities<br />

with "flow-through" systems (those that<br />

receive water from lentic or lotic environments).<br />

Eggs, other infective stages, and hosts can be<br />

carried with water that flows into and through<br />

facilities. Obviously this facilitates infection of<br />

fish. In general, trout in the present study being<br />

raised in flow-through systems had more parasite<br />

species and more individuals of the species<br />

present in comparison with the other systems.<br />

Similarly, Valtonen and Koskivaara (1994),<br />

studying the relationships between parasites of<br />

wild and cultured fishes in 2 lakes and a fish<br />

farm in Finland, reported that the source of parasites<br />

in the fish farm was the water-supplying<br />

lake.<br />

Cone and Cusack (1988) reported on the occurrence<br />

of 2 monogeneans, Gyrodactylus colemanensis<br />

Mizelle and Kritsky, 19<strong>67</strong>, and Gyrodactylus<br />

salmonis Yin and Sproston, 1948, on<br />

brook and rainbow trout, and Atlantic salmon,<br />

Salmo salar Linnaeus, 1758, in a farm in Nova<br />

Scotia, Canada, and discussed the origins of infection<br />

and their dispersal in the farm. Sources<br />

of infection with G. salmonis were stocks of infected<br />

rainbow trout brought into the facility<br />

from another farm, as well as wild infected Atlantic<br />

salmon and brook trout gaining access to<br />

the hatchery. Parasite dispersal in the farm involved<br />

infected fish jumping and wriggling from<br />

one pond to the next and the workers using<br />

transfer nets and buckets that contained live parasites.<br />

Also, brood stocks were infected and constituted<br />

internal reservoirs of infection.<br />

The effects of fish culture on natural waters<br />

receiving water from the farms has been a subject<br />

of increasing debate. If surveillance of parasites<br />

in the water above and below the fish facility<br />

is not continuous, little will be known<br />

about where the parasite really originated or<br />

how long it has been present. Regarding M. cerebralis,<br />

fish known to have whirling disease<br />

were imported into a commercial trout farm in<br />

Michigan in 1968. The receiving stream (a<br />

brook and brown trout stream) of this facility<br />

yielded M. cerefera/zs—infected rainbow trout escapees<br />

directly below the positive facility effluent.<br />

Valtonen and Koskivaara (1994) suggested<br />

that the farm itself was unlikely to affect the fish<br />

MUZZALL—PARASITES OF TROUT 187<br />

parasite fauna of the water-recipient lake, although<br />

some ectoparasites could originate from<br />

the farm.<br />

Muzzall (1995c), studying the parasites of<br />

pond-reared yellow perch, Perca flavescens<br />

(Mitchill, 1814) in Michigan, suggested that<br />

conditions of a pond associated with producing<br />

a good crop of fish also support a good crop of<br />

helminths that infect fish. Later, Muzzall (1996)<br />

referred to this as "the good fish crop-good helminth<br />

crop" relationship. Based on the results<br />

of the parasites infecting trout in the present<br />

study, this relationship does not occur. Of the<br />

parasite species found by Muzzall (1995c) infecting<br />

perch, 8 were represented as only larval<br />

stages, 6 of which were digenetic trematodes.<br />

Only 2 genera (generalist protozoans, Trichodina<br />

sp., Capriniana sp.) infesting perch were also<br />

found infesting trout. The dramatic differences<br />

in parasites found in yellow perch and trout from<br />

culture conditions can be explained by many<br />

factors. Probably the most important are the<br />

types of ponds used to culture the particular species,<br />

water temperatures, water sources, whether<br />

ponds are periodically drained, the surroundings<br />

of the ponds, and animals associated with the<br />

ponds.<br />

The state of control and prevention of parasites<br />

and diseases of fishes in culture in Michigan<br />

is difficult to assess. I refer to it as "crisis<br />

fisheries health," which can be defined as follows:<br />

"Some state and university officials, extension<br />

specialists, aquaculture centers, and trout<br />

farmers are not apparently concerned with fish<br />

health in aquaculture and in nature unless there<br />

is a crisis health problem, then action takes<br />

place." This approach is understandable with so<br />

many interested parties having different motives<br />

and the low priority of funding for parasite and<br />

disease work. I suggest that more studies on fish<br />

parasites and diseases in Michigan be encouraged<br />

and supported by the interested groups.<br />

Surveillance and surveys are needed to determine<br />

what parasites are infecting trout in culture<br />

conditions and in the surrounding waters.<br />

As mentioned earlier, growers in Michigan are<br />

involved in 3 activities in producing and selling<br />

trout. In regard to the first, the sale of infected<br />

trout for stocking could transfer some parasites<br />

to other fish directly or contaminate the watershed<br />

with other parasites. However, most if not<br />

all parasites reported in this study have been<br />

found infecting trout in the wild. Second, no par-<br />

Copyright © 2011, The Helminthological Society of Washington


188 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

asites were found that could infect humans if<br />

poorly cooked infected meat was eaten. Third,<br />

the sale of infected fish in fee-fishing ponds<br />

should not play a role in transmitting parasites,<br />

unless these fish are placed in other environments<br />

or the ponds have effluents to public waters.<br />

Obviously it should be emphasized that if<br />

trout are not routinely examined, light infections<br />

will not be noticed; when the infections do become<br />

evident, it may be too late to help the diseased<br />

fish.<br />

Acknowledgments<br />

I thank the trout farmers in Michigan who<br />

generously provided fish for this study, and Liz<br />

Osmer, Chris Henderson, Mindy Place, and Amy<br />

Hawkins for their technical assistance. I gratefully<br />

acknowledge Bob Baldwin, president of<br />

the Michigan Aquaculture Association, for making<br />

this study possible, and John Hnath for reviewing<br />

an early draft of the manuscript and<br />

sharing information with me on parasites.<br />

Literature Cited<br />

Allison, L. N. 1954. Advancements in prevention and<br />

treatment of parasitic diseases of fish. Transactions<br />

of the American Fisheries Society 83:221-228.<br />

, and W. C. Latta. 1969. Effects of gill lice<br />

(Salmincola edwardsii) on brook trout (Salvelinus<br />

fontinalis) in lakes. Michigan Department of Natural<br />

Resources, Research and Development Report<br />

No. 189. 32 pp.<br />

Anonymous. 1997. Michigan agricultural statistics<br />

1996—97. Trout. Commercial Fisheries Newsline<br />

(Michigan Sea Grant Extension)(December, 1997)<br />

16(2): 14.<br />

Boyce, N. P. J. 1969. Parasite fauna of pink salmon<br />

(Oncorhynchus gorbuscha) of the Bella Coola<br />

River, central British Columbia, during their early<br />

sea life. Journal of the Fisheries Research Board<br />

of Canada 26:813-820.<br />

Buchmann, K., and J. Bresciani. 1997. Parasitic infections<br />

in pond-reared rainbow trout Oncorhynchus<br />

mykiss in Denmark. Diseases of Aquatic Organisms<br />

28:125-138.<br />

Bullock, W. L. 1963. Intestinal histology of some salmonid<br />

fishes with particular reference to the histopathology<br />

of acanthocephalan infections. Journal<br />

of Morphology 112:23-44.<br />

Cone, D. K., and R. Cusack. 1988. A study of Gyrodactylus<br />

colemanensis Mizelle and Kritsky,<br />

19<strong>67</strong> and Gyrodactylus salmonis (Yin and Sproston,<br />

1948) (Monogenea) parasitizing captive salmonids<br />

in Nova Scotia. Canadian Journal of Zoology<br />

66:409-415.<br />

Hare, G. M., and C. Frantsi. 1974. Abundance and<br />

potential pathology of parasites infecting salmonids<br />

in Canadian Maritime hatcheries. Journal of<br />

Copyright © 2011, The Helminthological Society of Washington<br />

the Fisheries Research Board of Canada 31:1031-<br />

1036.<br />

Hernandez, A. D., and P. M. Muzzall. 1998. Seasonal<br />

patterns in the biology of Eubothrium salvelini<br />

infecting brook trout in a creek in lower<br />

Michigan. Journal of <strong>Parasitology</strong> 84:1119—1123.<br />

Hnath, J. G., ed. 1993. Great Lakes fish disease control<br />

policy and model program (supersedes September<br />

1985 edition). Great Lakes Fishery Commission<br />

Special Publication 93-1:1-38.<br />

Hoffman, G. L. 1984. Salmincola californiensis continues<br />

the march eastward. Fish Health Section,<br />

American Fisheries Newsletter 12:4.<br />

. 1999. Parasites of North American Freshwater<br />

Fishes. Cornell University Press, Ithaca, New<br />

York. 539 pp.<br />

Kabata, Z. 1969. Revision of the genus Salmincola<br />

Wilson, 1915 (Copepoda: Lernaeopodidae). Journal<br />

of the Fisheries Research Board of Canada 26:<br />

2987-3041.<br />

Lorn, J., and I. Dykova. 1992. Protozoan Parasites of<br />

Fishes. Developments in Aquaculture and Fisheries<br />

Science, 26. Elsevier Science Publishers, New<br />

York. 315 pp.<br />

Muzzall, P. M. 1984. Parasites of trout from four lotic<br />

localities in Michigan. Proceedings of the Helminthological<br />

Society of Washington 51:261-266.<br />

••—. 1986. Parasites of trout from the Au Sable<br />

River, Michigan, with emphasis on the population<br />

biology of Cystidicoloides tenuissima. Canadian<br />

Journal of Zoology 64:1549-1554.<br />

. 1993. Parasites of parr and lake age chinook<br />

salmon, Oncorhynchus tshawytscha, from the Pere<br />

Marquette River and vicinity, Michigan. Journal<br />

of the Helminthological Society of Washington<br />

60:55-61.<br />

. 1995a. Parasites of pacific salmon, Oncorhynchus<br />

spp., from the Great Lakes. Journal of Great<br />

Lakes Research 21:248-256.<br />

. 1995b. Parasites of lake trout, Salvelinus namaycush,<br />

from the Great Lakes: a review of the<br />

literature 1874-1994. Journal of Great Lakes Research<br />

21:594-598.<br />

. 1995c. Parasites of pond-reared yellow perch<br />

from Michigan. Progressive Fish-Culturist 57:<br />

168-172.<br />

. 1996. Parasites and diseases of pond-reared<br />

walleye and yellow perch. Aquaculture, November/December:49-61.<br />

Newman, J. R., and N. R. Kevern. 1994. Production<br />

of Michigan Aquacultural Products. Research Report<br />

526. Michigan Agricultural Experiment Station,<br />

Michigan <strong>State</strong> University, East Lansing. 78<br />

pp.<br />

Sawyer, T. K., J. G. Hnath, and J. F. Conrad. 1974.<br />

Thecamoeba hoffmani sp. n. (Amoebida: Thecamoebidae)<br />

from gills of fingerling salmonid fish.<br />

Journal of <strong>Parasitology</strong> 60:<strong>67</strong>7-682.<br />

Smith, H. D., and L. Margolis. 1970. Some effects<br />

of Eubothrium salvelini (Schrank, 1790) on sockeye<br />

salmon, Oncorhynchus nerka (Walbaum), in<br />

Babine Lake, British Columbia. Journal of <strong>Parasitology</strong><br />

56 (4, section 2, part l):321-322. (Abstract.)


Sutherland, D. R. 1999. Assessing the risk of whirling<br />

disease becoming established in the Great Lakes.<br />

Commercial Fisheries Newsline 18:10.<br />

, and D. D. Wittrock. 1985. The effects of<br />

Salmincola californiensis (Copepoda: Lernaeopodidae)<br />

on the gills of farm-raised rainbow trout,<br />

Salmo gairdneri. Canadian Journal of Zoology 63:<br />

2893-2901.<br />

MUZZALL—PARASITES OF TROUT 189<br />

Valtonen, E. T., and M. Koskivaara. 1994. Relationships<br />

between the parasites of some wild and culture<br />

fishes in two lakes and a fish farm in central<br />

Finland. International Journal for <strong>Parasitology</strong> 24:<br />

109-118.<br />

Yoder, W. G. 1972. The spread of Myxosoma cerebralis<br />

into native trout populations in Michigan.<br />

Progressive Fish-Culturist 34:103-106.<br />

Museums for Depositing of Specimens<br />

It is the policy of <strong>Comparative</strong> <strong>Parasitology</strong> to require the deposit of type and voucher specimens to document<br />

survey or taxonomic papers. Moreover, the value of any paper is enhanced by the deposit of reference specimens. The<br />

following museum collections in the United <strong>State</strong>s will accept such specimens, provide professional curatorial services<br />

for their preservation, provide accession numbers for inclusion in your publication, and make the deposited materials<br />

available for study by researchers worldwide. If other museum collections are used, they must provide comparable<br />

services, and information for contacting the museum must be provided in your publication. Materials designated as type<br />

specimens often carry the implication that they have been placed under such curatorial care; therefore, specimens retained<br />

in private collections should not carry type designations.<br />

The Editors would appreciate receiving additional information regarding other institutions that provide equivalent<br />

services. This information will be published in a future number of <strong>Comparative</strong> <strong>Parasitology</strong> and added to the directory<br />

that we are compiling for publication on the Society's website.<br />

Collection and domestic and international shipment of wildlife, including invertebrates, are governed by the laws<br />

and regulations of the countries of origination and destination. Detailed information for the United <strong>State</strong>s is available<br />

through the U.S. Fish and Wildlife Service website, http://www.fws.gov. It is also wise to contact the appropriate curator/<br />

collections manager of the receiving institution for special instructions before sending specimens for deposit.<br />

Helminths and Protozoans<br />

U.S. National Parasite Collection<br />

Biosystematics & National Parasite Collection Unit<br />

USDA, ARS, LPSI, BARC East No. 1180<br />

Beltsville, MD 20705-2350<br />

Curator: Dr. Eric P. Hoberg<br />

e-mail: ehoberg@lpsi.barc.usda.gov<br />

Telephone: (voice) 301-504-8444; (fax) -8979<br />

Homepage: http://www.lpsi.barc.usda.gov/bnpcu<br />

Harold W. Manter Laboratory of <strong>Parasitology</strong><br />

University of Nebraska <strong>State</strong> Museum<br />

W-529 Nebraska Hall,<br />

University of Nebraska<br />

Lincoln, NE 68588-0514<br />

Curator: Dr. Scott Lyell Gardner<br />

e-mail: slg@unl.edu<br />

Telephone: (voice) 402-472-3334; (fax) -8949<br />

Homepage: http://lamarck.unl.edu/research/parasitology<br />

Ticks<br />

U.S. National Tick Collection<br />

Institute of Arthropodology and <strong>Parasitology</strong><br />

Georgia Southern University<br />

<strong>State</strong>sboro, GA 30460-8056<br />

Curator: Dr. James E. Keirans<br />

e-mail: jkeirans@gasou.edu<br />

Telephone: (voice) 912-681-5564; (fax) -0559<br />

Homepage: http://www2.gasou.edu/iap/<br />

Mites<br />

J. Ralph Lichtenfels and Janet W. Reid<br />

National Mite Collection<br />

Systematic Entomology Laboratory<br />

USDA, ARS, BARC West No. 047<br />

Beltsville, MD 20705-2350<br />

Curator: Dug Miller<br />

e-mail: dmiller@sel.barc.usda.gov<br />

Telephone: (voice) 301-504-5895; (fax) -6842<br />

Homepage: http://www.sel.barc.usda.gov/Selhome/<br />

selhome.htm<br />

Crustaceans, Hirudineans<br />

Department of Invertebrate Zoology<br />

National Museum of Natural History<br />

Smithsonian Institution<br />

Washington, D.C. 20560-0163<br />

Collections Manager: Cheryl Bright<br />

e-mail: bright.cheryl@nmnh.si.edu<br />

Telephone: (voice) 202-357-4687; (fax) -3043<br />

Homepage: http://www.nmnh.si.edu/iz/collect.html<br />

Insects and Their Allies, Including Spiders<br />

Department of Entomology<br />

National Museum of Natural History<br />

Smithsonian Institution<br />

Washington, D.C. 20560-0165<br />

Collections Manager: Dr. David Furth<br />

e-mail: furth.david@nmnh.si.edu<br />

Telephone: (voice) 202-357-3146; (fax) 202-786-2894<br />

Homepage: http://entomology.si.edu<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 190-196<br />

Six New Host Records and an Updated List of Wild Hosts for<br />

Neobenedenia melleni (MacCallum) (Monogenea: Capsalidae)<br />

STEPHEN A. BuLLARD,1-5 GEORGE W. BENZ,2 ROBIN M. OVERSTREET,'<br />

ERNEST H. WILLIAMS, JR.,3 AND JAY HEMDAL4<br />

1 Gulf Coast Research Laboratory, Department of Coastal Sciences, University of Southern Mississippi, 703<br />

East Beach Drive, Ocean Springs, Mississippi 39564, U.S.A. (e-mail: ash.bullard@usm.edu;<br />

Robin.Overstreet@usm.edu),<br />

2 Tennessee Aquarium and Southeast Aquatic Research Institute, 1 Broad Street, RO. Box 1 1048,<br />

Chattanooga, Tennessee 37401, U.S.A. (e-mail: gwb@sari.org),<br />

3 Department of Marine Sciences, University of Puerto Rico, RO. Box 908, Lajas, Puerto Rico 006<strong>67</strong>-0908,<br />

U.S.A. (e-mail: bert@rmocfis.uprm.edu), and<br />

4 Toledo Zoo, 2700 Broadway, Toledo, Ohio 43609, U.S.A. (e-mail: jay.hemdal@toledozoo.org)<br />

ABSTRACT: Six new host records and an updated list of wild hosts for Neobenedenia melleni (MacCallum)<br />

(Monogenea: Capsalidae) are provided. We report specimens of N. melleni from the skin of a whitefin sharksucker<br />

(Echeneis neucratoides Zuieuw [Echeneidae]) caught off Mayagiiez, Puerto Rico; from the skin of a<br />

mosquitofish (Gambusia xanthosoma Greenfield [Poeciliidae]) caught in Little Salt Creek, Grand Cayman Island,<br />

British West Indies; from a freshwater immersion bath of red grouper (Epinephelus morio (Valenciennes) [Serranidae])<br />

caught in the Gulf of Mexico off Sarasota, Florida, U.S.A.; from the skin of a garden eel (Heteroconger<br />

hassi (Klausewitz and Eibl-Eibesfeldt) [Congridae]) in the Toledo Zoo, Toledo, Ohio, U.S.A.; from the skin of<br />

a raccoon butterflyfish (Chaetodon liinula (Cuvier) [Chaetodontidae]) in the Fort Wayne Children's Zoo, Fort<br />

Wayne, Indiana, U.S.A.; and from the gill cavity of a red snapper (Lutjanus campechanus (Poey) [Lutjanidae])<br />

in holding facilities at the Gulf Coast Research Laboratory, Ocean Springs, Mississippi, U.S.A. Neobenedenia<br />

melleni had not been reported previously from a suspected wild host in the Gulf of Mexico (i.e., E. morio) or<br />

from a member of Echeneidae, Atheriniformes, or Anguilliformes. Published host records indicate that N. melleni<br />

exhibits a relatively low degree of host specificity among captive and wild hosts; in nature, N. melleni infests<br />

predominantly shallow-water or reef teleosts.<br />

KEY WORDS: Neobenedenia melleni, Echeneis neucratoides, Gambusia xanthosoma, Epinephelus morio, Heteroconger<br />

hassi, Chaetodon lunula, Lutjanus campechanus, Monogenea, Capsalidae, host specificity, zoogeography,<br />

public aquaria, aquaculture, U.S.A., Puerto Rico, British West Indies, Florida, Mississippi, Gulf of Mexico.<br />

The capsalid Neobenedenia melleni (Mac- accounts of TV. melleni infesting wild hosts (see<br />

Callum, 1927) is relatively unusual among references in Table 1) are relatively scarce, and<br />

members of Monogenea in that it has been re- little is known about the breadth of host speciported<br />

from a wide range of hosts. This capsalid ficity exhibited by this parasite in nature. Thereinfests<br />

the eyes, fins, gill cavity, nasal cavity, fore, reports of TV. melleni from wild hosts are<br />

and skin of over 100 species of marine teleosts significant because they offer insight into the<br />

(Whittington and Horton, 1996). Most of these natural geographic distribution and host range of<br />

records are from fishes in aquaria and aquacul- this parasite. We report 6 new host records for<br />

ture systems where the parasite is identified as N. melleni: 3 from wild fishes and 3 from capa<br />

lethal pathogen (e.g., MacCallum, 1927; Jahn<br />

and Kuhn, 1932; Nigrelli and Breder, 1934;<br />

tive fishes,<br />

Mueller et al., 1994). However, there is no report Materials and Methods<br />

of disease associated with infestations of TV. me/leni<br />

among wild fishes. Neobenedenia melleni<br />

Worms were fixed in 10% neutral buffered formalin,<br />

70% ethanol or Bouin,s fixadve Eight worms were<br />

had been reported previously from wild hosts in stained in Van Cleave's hematoxylin containing sevthe<br />

Caribbean Sea, Gulf of California, and east- eral additional drops of Ehrlich's hematoxylin and<br />

ern Pacific Ocean off the coasts of Chile, Mex- were then dehydrated to 70% ethanol. Several drops<br />

ico, and<br />

j *u<br />

the<br />

TT<br />

United<br />

-4- j o«.<br />

<strong>State</strong>s<br />

« /T<br />

(Table<br />

ui i\<br />

1).<br />

ui-<br />

Published<br />

u j of aqueous<br />

. J' .<br />

saturated<br />

, ,<br />

lithium<br />

.<br />

carbonate<br />

. „ ,_.<br />

were<br />

,<br />

then<br />

.<br />

add-<br />

ed, followed by several drops of 6% butylamme<br />

.<br />

solution.<br />

Stained worms were dehydrated in an ethanol<br />

series, cleared in clove oil, and mounted permanently<br />

5 Corresponding author. on glass slides using neutral Canada balsam. Five<br />

190<br />

Copyright © 2011, The Helminthological Society of Washington


Table 1. Wild hosts for Neobenedenia melleni (MacCallum, 1927).<br />

Host Site Lo<br />

ATHERINIFORMES<br />

Poeciliidae<br />

Gambusia xanthosoma Greenfield. 1983<br />

Skin Little Salt Creek, Gra<br />

ish West Indies<br />

SCORPAENIFORMES<br />

Skin Southeast Pacific Oc<br />

Gills Northeast Pacific Oce<br />

Washington. U.S.A<br />

Mouth and skin Northeast Pacific Oc<br />

California, U.S.A.<br />

Scorpaenidae<br />

Sebastes capensis (Gmelin. 1789)<br />

Sebastes melanops Girard, 1856 (as Sebastodes melanops)<br />

Sebastes serranoides (Eigenmann and Eigenmann, 1890)<br />

Hexagrammidae<br />

Hexagrammos decagrammus (Pallas, 1810)<br />

Gills Northeast Pacific Oce<br />

Washington. U.S.A<br />

Cottidae<br />

Leptocottus armatus Girard, 1854<br />

Not indicated Northeast Pacific Oce<br />

nia, U.S.A.<br />

Gills Caribbean Sea off La<br />

PERCIFORMES<br />

Serranidae<br />

Epinephelus guttatus (Linnaeus, 1758)<br />

Not indicated Gulf of Mexico off S<br />

Epinephelus morio (Valenciennes, 1828)<br />

Not indicated Caribbean Sea off Bi<br />

Epinephelus striatus (Bloch, 1792)<br />

Gills Gulf of California of<br />

Mycteroperca rosacea (Gilbert, 1892) (as Mycteroperca pardalis)<br />

Skin Caribbean Sea off M<br />

Echeneidae<br />

Echeneis neucratoides Zuieuw. 1789<br />

Not indicated Caribbean Sea off Bi<br />

Lutjanidae<br />

Lutjanus apodus (Walbaum, 1892) (as Lutianus apodus)<br />

Copyright © 2011, The Helminthological Society of Washington


Table 1. Continued.<br />

Host Site Loca<br />

Sparidae<br />

Archosargus probatocephalus (Walbaum 1792)j<br />

Northeast Pacific Ocea<br />

nia, U.S.A.<br />

Eyes and skin<br />

Caribbean Sea off Bim<br />

Not indicated<br />

Chaetodontidae<br />

Chaetodon capistratus Linnaeus, 1758<br />

Caribbean Sea off Bim<br />

Not indicated<br />

Chaetodon ocellatits Bloch, 1787<br />

Caribbean Sea off Bim<br />

Not indicated<br />

Chaetodon striatus Linnaeus, 1758<br />

Caribbean Sea off Bim<br />

Not indicated<br />

Pomacanthidae<br />

Holocanthus ciliaris (Linnaeus, 1758)<br />

Caribbean Sea off Bim<br />

Not indicated<br />

Holocanthus tricolor (Bloch, 1795)<br />

Caribbean Sea off Bim<br />

Not indicated<br />

Pomacanthus arcuatux (Linnaeus, 1758)<br />

Caribbean Sea off Bim<br />

Not indicated<br />

Pomacanthus paru (Bloch, 1787)<br />

Skin<br />

Kyphosidae<br />

Cirella nigricans (Ayres, 1860)<br />

Northeast Pacific Ocean<br />

nia, U.SA.<br />

Northeast Pacific Ocean<br />

land, California, U.S<br />

Northeast Pacific Ocean<br />

land, California, U.S<br />

Fins and skin<br />

Skin<br />

Medialuna californiensis (Steindachner, 1876)<br />

Exterior<br />

Embiotocidae<br />

Embiotoca jacksoni Agassiz, 1853<br />

Northeast Pacific Ocean<br />

ta Barbara, Californi<br />

Northeast Pacific Ocean<br />

California, U.S.A.<br />

Northeast Pacific Ocean<br />

California, U.S.A.<br />

Northeast Pacific Ocean<br />

California, U.S.A.<br />

Exterior of head<br />

Embiotoca lateralis Agassiz, 1854<br />

Exterior<br />

Exterior<br />

Rhacochilus vacca (Girard, 1855) (as Damalichthys vacca)<br />

Copyright © 2011, The Helminthological Society of Washington


CB <<br />

•£ a<br />

o '=<br />

2 I<br />

BULLARD ET AL.—HOSTS OF NEOBENEDENIA MELLENl 193<br />

worms intended for study using Nomarski illumination<br />

were dehydrated, cleared in clove oil, and mounted<br />

unstained in neutral Canada balsam. Worms were identified<br />

using the original description of N. melleni (as<br />

Epibdella melleni MacCallum, 1927), the redescription<br />

of N. melleni contained in a recent revision of Neobenedenia<br />

Yamaguti, 1963, and the key to the species<br />

of Neobenedenia (see Whittington and Horton, 1996).<br />

We primarily used 1) anterior attachment organs circular<br />

and not bipartite; 2) anterior hamuli recurved,<br />

nonserrated (i.e., smooth), and robust (i.e., width usually<br />

greater than that of both accessory sclerites and<br />

posterior hamuli and with root of consistent width<br />

along total length [i.e., root not tapered, constricted, or<br />

pinched]); 3) glands of Goto not evident, and 4) other<br />

specific features indicated by Whittington and Horton<br />

(1996). Nomenclature used herein for members of<br />

Neobenedenia follows that of Whittington and Horton<br />

(1996). Specimens of N. melleni from Gambusia xanthosoma<br />

(Poeciliidae) and Lutjanus campechanus<br />

(Poey, 1860) (Lutjanidae) were deposited in the United<br />

<strong>State</strong>s National Parasite Collection (USNPC) at Beltsville,<br />

Maryland, U.S.A. (USNPC Nos. 089159 and<br />

089160), and specimens from Echeneis neucratoides<br />

(Echeneidae), Epinephelus morio (Serranidae), Heteroconger<br />

hassi (Klausewitz and Eibl-Eibesfeldt, 1959)<br />

(Congridae), and Chaetodon lunula (Cuvier, 1831)<br />

(Chaetodontidae) were deposited there (USNPC Nos.<br />

089161, 089162, 089163, and 089164) and in the helminth<br />

collections of the H. W. Manter Laboratory<br />

(HWML) of the University of Nebraska <strong>State</strong> Museum<br />

at Lincoln, Nebraska, U.S.A. (HWML Nos. 15063,<br />

15064, 15065, and 15066).<br />

Results and Discussion<br />

Regarding our new host records, 2 specimens<br />

of N. melleni were collected from the skin of a<br />

whitefin sharksucker (E. neucratoides), a rernora<br />

that was attached to a West Indian manatee (Trichechus<br />

manatus Linnaeus, 1758 [Trichechidae])<br />

off Mayagiiez, Puerto Rico. This is the first report<br />

of N. melleni from a remora and may help<br />

to explain in part the wide geographic distribution<br />

of N. melleni. Although carriers of infested<br />

remoras may not travel between oceans, infested<br />

remoras may transfer infestations of N. melleni<br />

among fish, mammalian, and turtle species and<br />

individuals with which they associate. In addition,<br />

remoras can attached to or mingle with<br />

their carriers for prolonged periods of time. This<br />

habit may provide N. melleni opportunity to infest<br />

the remora's carrier host or other fishes in<br />

close proximity to the infested remora. Various<br />

cleaner fishes (e.g., bluehead wrasse, Thalassoma<br />

bifasciatum (Bloch, 1791) [Labridae]; neon<br />

goby, Gobiosoma oceanops (Jordan, 1904) [Gobiidae];<br />

and cleaning goby, Gobiosoma genie<br />

Bohlke and Robins, 1968) were effective in controlling<br />

infestations of N. melleni among aquar-<br />

Copyright © 2011, The Helminthological Society of Washington


194 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

ium-kept fish (see Cowell et al., 1993). Some<br />

species of remora feed on ectoparasites (Cressey<br />

and Lachner, 1970), and, because of this, aquaculturists<br />

eventually may use remoras to control<br />

infestations of N. melleni on large hosts. However,<br />

as previously suggested, remoras may<br />

transport worms to adjacent groups of fishes.<br />

A specimen of TV. melleni was collected from<br />

the skin of a mosquitofish (G. xanthosoma) from<br />

Little Salt Creek (western shore of North Sound,<br />

Grand Cayman, British West Indies). Neobenedenia<br />

melleni had not been reported previously<br />

from a member of Atheriniformes or from the<br />

western Caribbean Sea. The specimen of N. melleni<br />

was conspicuous, 3 mm in total length, and<br />

attached to the dorsal surface of the head at the<br />

level of the eyes of a mosquitofish that was 33<br />

mm in total length. Gambusia xanthosoma is apparently<br />

endemic to the high salinity mangrove<br />

habitats throughout North Sound (Abney and<br />

Heard, personal communication); therefore, it is<br />

of ecological interest to report on the occurrence<br />

of nonendemic parasites, such as N. melleni, that<br />

infest a wide range of hosts and that are identified<br />

as lethal pathogens among confined fishes.<br />

Nigrelli (1947) reported several wild hosts for<br />

N. melleni in the Caribbean Sea off Bimini (see<br />

Table 1). Robinson et al. (1992) and Hall (1992)<br />

reported heavy infestations of N. melleni among<br />

cultured, seawater-acclimated red hybrid tilapia<br />

in floating cages off southern Jamaica. Cowell<br />

et al. (1993) reported infestations of TV. melleni<br />

on Florida red tilapia (descendants of an original<br />

cross between Oreochromis urolepis hornorum<br />

(Norman, 1922) [Cichlidae] and Oreochromis<br />

mossambicus (Peters, 1852)) in aquaria at the<br />

Caribbean Marine Research Center (CMRC),<br />

Lee Stocking Island, Exuma Cays, Bahamas.<br />

However, because N. melleni has a broad host<br />

range and wide geographic distribution and<br />

heavily infests some hosts in aquaculture, we<br />

cannot determine if, when, or how it was introduced<br />

to the endemic population of G. xanthosoma.<br />

At least 3 specimens of TV. melleni infested<br />

the red grouper (E. morio); they were caught off<br />

Sarasota, Florida, U.S.A., in January 1993. Material<br />

of TV. melleni was collected from a freshwater<br />

immersion bath at the Mote Marine Laboratory<br />

(MML), Sarasota, Florida, when the fish<br />

were initially treated after being captured from<br />

the Gulf of Mexico. Nevertheless, TV. melleni later<br />

became established in culture facilities at the<br />

Copyright © 2011, The Helminthological Society of Washington<br />

MML. Neobenedenia melleni was previously reported<br />

from E. morio and Mycteroperca tnicrolepis<br />

(Goode and Bean, 1879) (Serranidae) in<br />

recirculating-seawater tanks in northwestern<br />

Florida (Florida <strong>State</strong> University Marine Laboratory,<br />

Turkey Point, Florida, U.S.A.) by Mueller<br />

et al. (1994) and from other members of the<br />

sea bass family in the Caribbean Sea and the<br />

Gulf of California (see Table 1); however, this<br />

is the first report of TV. melleni from a suspected<br />

wild host in the Gulf of Mexico.<br />

We also report numerous adult and juvenile<br />

specimens of TV. melleni from the skin of a garden<br />

eel (H. hassi) from the Toledo Zoo, Toledo,<br />

Ohio, U.S.A. This is the first report of TV. melleni<br />

from any member of Anguilliformes, and to the<br />

best of our knowledge, it is also the first report<br />

of TV. melleni from a host that lacks scales.<br />

Whereas the exact geographic origin of the eel<br />

was not known, we suspect that it became infested<br />

while confined in a compartmentalized<br />

quarantine system at the Toledo Zoo. One of us<br />

(J.H.) observed a cream angelfish (Apolemichthys<br />

xanthurus (Bennett, 1832) [Pomacanthidae])<br />

in this same water system that harbored<br />

numerous specimens of a platyhelminth on its<br />

skin that were probably TV. melleni. Nigrelli and<br />

Breder (1934) reported that some angelfishes<br />

were foci for epidemics of TV. melleni in the New<br />

York Aquarium. Specimens of TV. melleni have<br />

yet to be reported from A. xanthurus. However,<br />

because the aforementioned worms from this<br />

host were not available for identification, we did<br />

not report this fish as a host for TV. melleni.<br />

Numerous specimens of TV. melleni were also<br />

collected from a raccoon butterflyfish (C. lunula)<br />

that died while in quarantine at the Fort<br />

Wayne Children's Zoo, Fort Wayne, Indiana,<br />

U.S.A. We are not certain of the exact geographic<br />

origin of that wild-caught fish or whether it<br />

was infested in the wild. However, C. lunula is<br />

a reef species that ranges from East Africa to<br />

Polynesia (Randall et al., 1990), and that raccoon<br />

butterflyfish most likely came from there<br />

(i.e., Indo-Pacific Region). Neobenedenia melleni<br />

has been reported from 3 members of Chaetodon<br />

in the Caribbean Sea (Nigrelli, 1947; Table<br />

1).<br />

A single specimen of TV. melleni was collected<br />

from the gill cavity of a red snapper (L. campechanus)<br />

caught in the northern Gulf of Mexico<br />

and maintained in an aquaculture tank with<br />

other red snapper at the Gulf Coast Research


Laboratory (GCRL), Ocean Springs, Mississippi,<br />

U.S.A. The tank and filtration system that<br />

supported this host had been sanitized before<br />

adding any fish, and no other fishes shared the<br />

water of this system. There was no history of<br />

infestation by this monogenean in culture facilities<br />

at GCRL. Therefore, it is likely that this red<br />

snapper was infested with N. melleni in the wild.<br />

In addition, 3 juvenile red snapper (each approximately<br />

120 mm in total length) that were<br />

spawned and reared at the GCRL aquaculture<br />

facility and then transferred to the GCRL Marine<br />

Education Center (MEC), Biloxi, Mississippi,<br />

U.S.A., became heavily infested with TV. melleni.<br />

These red snapper were maintained in a 7,700liter<br />

aquarium with a spadefish (Chaetodipterus<br />

faber (Broussonet, 1782) [Ephippidae]) that was<br />

also heavily infested with the monogenean. Nigrelli<br />

(1947) reported Lutjanus apodus (as Lutianus<br />

apodus) as a wild host for N. melleni (as<br />

Benedenia melleni) in the West Indies. We suspect<br />

that L. campechanus also may be a wild<br />

host of N. melleni in the northern Gulf of Mexico.<br />

However, it does not seem to be a common<br />

host, because we have yet to observe a specimen<br />

of N. melleni on a red snapper directly from the<br />

wild, in spite of examinations of at least 276<br />

such fish.<br />

The most recent list of captive and wild hosts<br />

for N. melleni was presented by Lawler (1981).<br />

Whittington and Horton (1996) subsequently<br />

provided a list of hosts for N. melleni; however,<br />

that list did not distinguish between captive and<br />

wild hosts. Because a list identifying wild hosts<br />

for N. melleni has not been presented in 18<br />

years, we consider Table 1 a useful update.<br />

Rarely does a monogenean species, let alone<br />

a capsalid, occur in more than 1 ocean and infest<br />

more than 1 host species, and if so, those hosts<br />

are usually closely related species (e.g., Byrnes<br />

and Rohde, 1992; Whittington, 1998). Neobenedenia<br />

melleni has now been reported from 27<br />

species comprising 18 genera, 14 families, and<br />

3 orders of wild hosts (Table 1). These records<br />

suggest that N. melleni is a parasite of predominantly<br />

shallow-water or reef-dwelling marine<br />

teleosts. Neobenedenia melleni exhibits a relatively<br />

low degree of host specificity among both<br />

captive and wild hosts. Nigrelli and Breder<br />

(1934) studied the host-parasite relationship between<br />

N. melleni and several fishes held at the<br />

New York Aquarium. However, the factors that<br />

allowed it to infest a broad array of hosts in<br />

BULLARD ET AL.—HOSTS OF NEOBENEDENIA MELLENI 195<br />

captivity and in the wild were not clearly understood.<br />

In some cases, horizontal transfer and<br />

levels of infestation may be limited initially only<br />

by the physical distance between parasite and<br />

potential host. This could, in part, explain the<br />

apparent abundance of N. melleni among reef<br />

fishes that live in close proximity to one another<br />

in the wild and among those and other fishes<br />

held in public aquaria and aquaculture systems.<br />

Further study of this unique monogenean utilizing<br />

molecular techniques could possibly reveal<br />

population differences.<br />

Acknowledgments<br />

We thank Reg Blaylock for commenting on<br />

the manuscript; Nate Jordan, Jason Sleekier,<br />

Jody Peterson, and Casey Nicholson (all of<br />

GCRL) for providing red snapper for examination;<br />

Joyce Shaw (GCRL) for requesting some<br />

of the pertinent literature via interlibrary loan;<br />

Alex Schesny (MEC) for providing juvenile<br />

specimens of L. campechanus infested with N.<br />

melleni; Michael Abney (University of Kentucky)<br />

and Richard Heard (GCRL) for providing<br />

the specimen of G. xanthosoma infested with TV.<br />

melleni; Pamela Phelps (MML) for providing<br />

specimens of N. melleni from E. morio; David<br />

Miller (Fort Wayne Children's Zoo) for providing<br />

specimens of N. melleni from C. lunula; and<br />

the Cayman Islands National Trust and the Cayman<br />

Islands Department of the Environment for<br />

allowing and facilitating collection of G. xanthosoma<br />

on Grand Cayman. This study was supported<br />

in part from National Oceanic and Atmospheric<br />

Administration, National Marine<br />

Fisheries Service, award No. NA86FL0476 and<br />

NA96FL0358.<br />

Literature Cited<br />

Bravo-Hollis, M. 1957. Trematodos de peces marinos<br />

de aguas mexicanas. XIV. Cuatro monogeneos de<br />

la familia Capsalidae Baird, 1853, de las costas<br />

del Pacifico, incluyendo una especie nueva. Anales<br />

del Institute de Biologia, Mexico 28:195-216.<br />

, and J. C. Deloya. 1973. Catalogo de la coleccion<br />

helmintologica del Institute de Biologia.<br />

Institute de Biologia, Universidad Nacional Autonoma<br />

de Mexico, Publicaciones Especiales 2.<br />

137 pp.<br />

Brooks, D. R., and M. A. Mayes. 1975. Phyllodistomum<br />

scrippsi sp. n. (Digenea: Gorgoderidae)<br />

and Neobenedenia girellae (Hargis, 1955) Yamaguti,<br />

1963 (Monogenea: Capsalidae) from the California<br />

sheephead, Pimelometopon pulchrum (Ayers)<br />

(Pisces: Labridae). Journal of <strong>Parasitology</strong> 61:<br />

407-408.<br />

Copyright © 2011, The Helminthological Society of Washington


196 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Byrnes, T., and K. Rohde. 1992. Geographical distribution<br />

and host specificity of ectoparasites of<br />

Australian bream, Acanthopagrus spp. (Sparidae).<br />

Folia Parasitologica 39:249-264.<br />

Cowell, L. E., W. O. Watanabe, W. D. Head, J. J.<br />

Grover, and J. M. Shenker. 1993. Use of tropical<br />

cleaner fish to control the parasite Neobenedenia<br />

melleni (Monogenea: Capsalidae) on seawater-cultured<br />

Florida red tilapia. Aquaculture<br />

113:189-200.<br />

Cressey, R. F., and E. A. Lachner. 1970. The parasitic<br />

copepod diet and life history of diskfishes<br />

(Echeneidae). Copeia 1970:310-318.<br />

Dyer, W. G., E. H. Williams, and L. Bunkley-Williams.<br />

1992. Neobenedenia pargueraensis n. sp.<br />

(Monogenea: Capsalidae) from the red hind, Epinephelus<br />

giittatus, and comments about Neobenedenia<br />

melleni. Journal of <strong>Parasitology</strong> 78:399-<br />

401.<br />

Gaida, I. H., and P. Frost. 1991. Intensity of Neobenedenia<br />

girellae (Monogenea: Capsalidae) on<br />

the halfmoon, Mediahma californiensis (Perciformes:<br />

Kyphosidae), examined using a new method<br />

for detection. Journal of the Helminthological Society<br />

of Washington 58:129-130.<br />

Goldberg, J. L., R. Millar, and S. Sanchez. 1991.<br />

The ontogenic acquisition of infestation of the<br />

trematode ectoparasite Neobenedenia girellae on<br />

the marine teleost Girella nigricans. Bulletin of<br />

the Southern California Academy of Sciences 90:<br />

83-85.<br />

Gonzalez, M. T., and E. Acuna. 1998. Metazoan parasites<br />

of the red rockfish Sebastes capensis off<br />

northern Chile. Journal of <strong>Parasitology</strong> 84:783-<br />

788.<br />

Hall, R. N. 1992. Preliminary investigations of marine<br />

cage culture of red hybrid tilapia in Jamaica. Proceedings<br />

of the Gulf and Caribbean Fisheries Institute<br />

42:440.<br />

Hargis, W. J. 1955. A new species of Benedenia<br />

(Trematoda: Monogenea) from Girella nigricans,<br />

the opaleye. Journal of <strong>Parasitology</strong> 41:48-50.<br />

Jahn, T. L., and L. R. Kuhn. 1932. The life history<br />

of Epibdella melleni MacCallum 1927, a monogenetic<br />

trematode parasitic on marine fishes. Biological<br />

Bulletin 62:89-111.<br />

Lawler, A. R. 1981. Zoogeography and host-specificity<br />

of the superfamily Capsaloidea Price, 1936<br />

(Monogenea: Monopisthocotylea): an evaluation<br />

of the host-parasite locality records of the superfamily<br />

Capsaloidea Price, 1936, and their utility<br />

Copyright © 2011, The Helminthological Society of Washington<br />

in determinations of host-specificity and zoogeography.<br />

Special Papers in Marine Science No. 6<br />

(Virginia Institute of Marine Science, Gloucester<br />

Point, Virginia, U.S.A.). 650 pp.<br />

Love, M. S., K. Shriner, and P. Morris. 1984. Parasites<br />

of olive rockfish, Sebastes serranoides,<br />

(Scorpaenidae) off central California. Fishery Bulletin<br />

82:530-537.<br />

MacCallum, G. A. 1927. A new ectoparasitic trematode.<br />

Zoopathologica 1:291-300.<br />

Moser, M., and L. Haldorson. 1982. Parasites of two<br />

species of surfperch (Embiotocidae) from seven<br />

Pacific coast locales. Journal of <strong>Parasitology</strong> 68:<br />

733-735.<br />

Mueller, K. W., W. O. Watanabe, and W. D. Head.<br />

1994. Occurrence and control of Neobenedenia<br />

melleni (Monogenea: Capsalidae) in cultured tropical<br />

marine fish, including three new host records.<br />

Progressive Fish-Culturist 56:140—142.<br />

Nigrelli, R. F. 1947. Susceptibility and immunity of<br />

marine fishes to Benedenia ( = Epibdella) melleni<br />

(MacCallum), a monogenetic trematode. III. Natural<br />

hosts in the West Indies. Journal of <strong>Parasitology</strong><br />

33:25.<br />

, and C. M. Breder, Jr. 1934. The susceptibility<br />

and immunity of certain marine fishes to<br />

Epibdella melleni, a monogenetic trematode. Journal<br />

of <strong>Parasitology</strong> 20:259-269.<br />

Randall, J. E., G. R. Allen, and R. C. Steene. 1990.<br />

Fishes of the Great Barrier Reef and Coral Sea.<br />

University of Hawaii Press, Honolulu, Hawaii.<br />

507 pp.<br />

Robins, C. R., G. G. Ray, and J. Douglas. 1986. A<br />

Field Guide to Atlantic Coast Fishes of North<br />

America. Houghton Mifflin Company, Boston,<br />

Massachusetts. 354 pp.<br />

Robinson, R. D., L. F. Khalil, R. N. Hall, and R. D.<br />

Steele. 1992. Infection of red hybrid tilapia with<br />

a monogenean in coastal waters off southern Jamaica.<br />

Proceedings of the 42nd Annual Gulf and<br />

Caribbean Fisheries Institute 42:441-447.<br />

Whittington, I. D. 1998. Diversity "down under":<br />

monogeneans in the antipodes (Australia) with a<br />

prediction of monogenean biodiversity worldwide.<br />

International Journal for <strong>Parasitology</strong> 28:1481-<br />

1493.<br />

, and M. A. Horton. 1996. A revision of Neobenedenia<br />

Yamaguti, 1963 (Monogenea: Capsalidae)<br />

including a redescription of N. melleni<br />

(MacCallum, 1927) Yamaguti, 1963. Journal of<br />

Natural History 30:1113-1156.


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 197-201<br />

Hymenolepis nana in Pet Store Rodents<br />

LAURA M. DUCLOS AND DENNIS J. RICHARDSON'<br />

Department of Biological Sciences, Box 138, 275 Mount Carmel Avenue, Quinnipiac University, Hamden,<br />

Connecticut 06518, U.S.A. (e-mail: dennis.richardson@quinnipiac.edu)<br />

ABSTRACT: The rodent tapeworm, Hymenolepis nana, is a zoonotic pathogen transmissible through the ingestion<br />

of eggs in feces or cysticercoids in arthropods. Since data addressing the potential for acquiring human infections<br />

of H. nana from pet rodents are lacking, a survey of pet stores in southern Connecticut, U.S.A., was conducted.<br />

Fecal flotation analysis revealed 9.1% overall prevalence in 110 samples collected weekly from cages holding<br />

group-housed small animals, from 3 stores for 4 weeks. Of 11 species, only cages holding rats (3 of 22 samples),<br />

mice (6 of 30 samples), and prairie dogs (1 of 2 samples) were positive. Necropsies of 38 rats, 72 domestic<br />

mice, and 39 golden hamsters purchased from 9 stores showed prevalences of 31.6%, 22.2%, and 10.3%,<br />

respectively. Mean intensity was 66 worms per rat, 14 worms per mouse, and 15 worms per hamster. Overall,<br />

75% of surveyed pet stores were selling infected rats, mice, or hamsters, indicating that pet store rodents pose<br />

a potential threat to public health.<br />

KEY WORDS: Hymenolepis nana, pets, rodents, survey, zoonosis.<br />

Hymenolepis nana Siebold, 1852, infects 75<br />

million people worldwide (Crompton, 1999), of<br />

whom the majority are children (Little, 1985;<br />

Markell et al., 1999). Hymenolepis nana has a<br />

cosmopolitan distribution, with human, Old<br />

World monkey, and rodent definitive hosts becoming<br />

infected through ingestion of infective<br />

cysticercoids within beetle or flea intermediate<br />

hosts or through ingestion of eggs in feces. In<br />

the latter route, cysticercoids develop in intestinal<br />

villi, with worms later reemerging and attaching<br />

to the mucosa (Roberts and Janovy,<br />

<strong>2000</strong>). Direct transmission through the ingestion<br />

of eggs may be the most common route of infection<br />

in humans (Turner, 1975).<br />

It has been suggested that 2 morphologically<br />

identical subspecies of H. nana exist, yet this<br />

tapeworm is typically classified as a zoonotic<br />

and is capable of horizontal transmission between<br />

human and nonhuman animals (Fox et al.,<br />

1984; Jacoby and Fox, 1984; Chomel, 1992).<br />

Human infections "produce either no symptoms<br />

or vague abdominal disturbances. In fairly heavy<br />

infections, children may show lack of appetite,<br />

abdominal pain with or without diarrhea, anorexia,<br />

vomiting, and dizziness" (Neva and<br />

Brown, 1994). Such nondescript symptoms may<br />

account for the low number of reported clinical<br />

cases, and many subclinical infections may go<br />

undiagnosed. Available prevalence data regarding<br />

human populations were obtained in most<br />

cases from fecal surveys conducted in develop-<br />

1 Corresponding author.<br />

197<br />

ing nations. For example, H. nana was found in<br />

20.5% of Australian aborigines (Meloni et al.,<br />

1993), 8-10% of oncology patients in Mexico<br />

(Guarner et al., 1997), 16% of Egyptian school<br />

children (Khalil et al., 1991), and 0.6% of Thai<br />

laborers (Wilairatana et al., 1996). In developed<br />

regions such as Western Europe and North<br />

America, human infections are seldom identified<br />

or acknowledged, and survey data are often<br />

patchy and scarce (Croll and Gyorkos, 1979; Jacobs,<br />

1979; Seaton, 1979; Cooper et al., 1981).<br />

Still, H. nana is estimated to be an important<br />

cause of cestodiasis in the southeastern United<br />

<strong>State</strong>s, with infections found in approximately<br />

1% of school children (Roberts and Janovy,<br />

<strong>2000</strong>) and 4% of pediatric clinic patients (Flores<br />

et al., 1983). In 1987, 34 state diagnostic laboratories<br />

identified H. nana in collected stool<br />

samples (0.4%), with Connecticut reporting a<br />

prevalence of 0.8%, Massachusetts 0.4%, and<br />

Rhode Island 1.6% (Kappus et al., 1991).<br />

Most studies focus on human infection, but<br />

fail to adequately address potential zoonotic<br />

sources of the infection. In Turkey, 5.6% of surveyed<br />

wild mice and rats harbored H. nana (Sahin,<br />

1979), and in Saudi Arabia, H. nana was<br />

reported from baboons living in close proximity<br />

to humans (Ghandour et al., 1995). In the United<br />

<strong>State</strong>s, Stone and Manwell (1966) reported infection<br />

in 21% of mice and 9% of hamsters from<br />

Syracuse University, Syracuse, New York,<br />

U.S.A., animal rooms and various commercial<br />

vendors. In the same study, pet mice and ham-<br />

Copyright © 2011, The Helminthological Society of Washington


198 uuMKAKAiIVt, rAKAsuOLOuY, 6/(2), JULY <strong>2000</strong><br />

Table 1. Results of cage sampling for Hymenolepis nana using fecal flotation.<br />

Host species<br />

Domestic spiny mouse (Heteromyidae)<br />

Long-tailed chinchilla (Chinchilla lanigcra Molina, 1782)<br />

Black-tailed prairie dog (Cynomys ludovicianus Ord, 1815)<br />

Guinea pig (Cavia porcellus Linnaeus, 1758)<br />

Domestic mouse (Mus tnusculus Linnaeus, 1758)<br />

Ferret (Mustelaputoriusfa.ro Linnaeus, 1758)<br />

Mongolian gerbil (Meriones unguiculatus Milne-Edwards, 18<strong>67</strong>)<br />

European rabbit (Oryctolagux cuniculus Linnaeus, 1758)<br />

Siberian hamster (Phodopus sungorus Pallus, 1773)<br />

Norway rat (Rattus norvegicus Berkenhout, 1769)<br />

:i: Overall prevalence of infected cages was 9.1%.<br />

t A total of 5 individual prairie dogs was surveyed from the 2 cage samples.<br />

:i: Hymenolepis dirninuta eggs were also detected in the cage sample.<br />

sters showed prevalences of 66% and 44% respectively.<br />

Pet stores are traditionally implicated as potential<br />

sources of human parasite infections, but<br />

emphasis is centered upon feline, canine, or avian<br />

species rather than rodents. However, H.<br />

nana is a common zoonosis of pet rodents (Chomel,<br />

1992), and in 1969 infection was detected<br />

in Mongolian gerbils purchased as pets from a<br />

department store (Lussier and Loew, 1970). Given<br />

that children have less than optimal hygiene<br />

habits, and immune-compromised individuals,<br />

such as those with the acquired immunodeficiency<br />

syndrome (AIDS) or undergoing cancer<br />

treatment, are at greater risk for disease (Gerba<br />

et al., 1996), pet rodent infections raise obvious<br />

public health concerns. Additionally, there is a<br />

lack of survey data addressing the assumption<br />

that golden hamsters are more often parasitized<br />

with H. nana than are other rodents (Chomel,<br />

1992; Teclaw et al., 1992). The purpose of this<br />

study was to assess health risks associated with<br />

human and rodent interaction as they pertain to<br />

H. nana, through a survey of small animals sold<br />

by pet stores in southern Connecticut.<br />

Materials and Methods<br />

Once a week for 4 weeks beginning in July 1999, a<br />

fecal survey was conducted on all small animal cages<br />

from 3 pet stores. Samples of 5-10 fecal pellets were<br />

collected from the bedding of cages housing grouped<br />

animals and analyzed by fecal flotation (Hendrix,<br />

1998). A total of 110 cage samples was obtained from<br />

representatives of 11 domesticated small animal species<br />

(Table 1).<br />

Based on the findings from fecal analysis of small<br />

No. of cage<br />

samples<br />

3<br />

6<br />

2<br />

19<br />

18<br />

30<br />

3<br />

7<br />

11<br />

1<br />

22$<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Samples ( + )<br />

for H. nana<br />

0<br />

0<br />

1<br />

0<br />

0<br />

6<br />

0<br />

0<br />

0<br />

0<br />

3<br />

Cage<br />

prevalence<br />

(%)*<br />

0<br />

0<br />

sot<br />

0<br />

0<br />

20<br />

0<br />

0<br />

0<br />

0<br />

14<br />

animal cages, individual rodents were purchased from<br />

9 different pet stores not included in the fecal survey,<br />

and a postmortem examination of the intestinal tract<br />

of each rodent was performed. Necropsy was conducted<br />

on a total of 38 rats, 39 golden hamsters, and<br />

72 domestic mice. Animals were killed by CO2 narcosis,<br />

and the small intestine, from the pyloric sphincter<br />

to the ileocecal juncture, was removed, placed in a<br />

Petri dish of tap water, and opened longitudinally.<br />

Worms were removed and counted. Representative<br />

specimens were stained, mounted, and deposited in the<br />

United <strong>State</strong>s National Museum Parasite Collection in<br />

Beltsville, Maryland, U.S.A. (USNPC No. 089330.00).<br />

Results<br />

Fecal analysis showed that 9.1% of cages<br />

housed infected animals, with animals from 1<br />

pet store testing positive for 3 of 4 weeks. Domestic<br />

mice and Norway rats exhibited prevalences<br />

of 30.0% and 13.6%, respectively. One<br />

of 2 black-tailed prairie dog cage samples revealed<br />

H. nana. All other species, including<br />

golden hamsters, were negative by fecal flotation<br />

analysis (Table 1).<br />

Necropsy results of purchased animals revealed<br />

that 7 of 9 pet stores were selling infected<br />

rats, domestic mice, and/or golden hamsters.<br />

Prevalence was highest in rats (31.6%), with<br />

mean intensity (MI) of 66 worms per host.<br />

Mouse prevalence was lower at 22.2% (MI =<br />

15), and only 4 golden hamsters (10.3%) were<br />

infected (MI = 15). One rat was infected with<br />

Hymenolepis dirninuta Rudolphi, 1819 (Table 2).<br />

Discussion<br />

Rodents typically remained in pet stores approximately<br />

7-10 days. The prepatent period for


Table 2. Presence of Hymenolepis nana in necropsied<br />

rats, mice, and hamsters.<br />

No. of No. (%) of<br />

individuals individuals Mean<br />

Host species necropsied infected intensity<br />

Golden hamster<br />

Domestic mouse<br />

Norway rat<br />

39<br />

72<br />

38<br />

4 (10.3)<br />

16 (22.2)<br />

12 (31.6)*<br />

* One rat was infected with Hymenolepis dimimita.<br />

DLJCLOS AND RICHARDSON—HYMENOLEPIS NANA IN PETS 199<br />

15<br />

14<br />

66<br />

H. nana is approximately 25 days (Hunninen,<br />

1935; Jacoby and Fox, 1984), leading to the conclusion<br />

that animals arrived infected from commercial<br />

vendors or private breeders, rather than<br />

acquiring infection through exposure at pet<br />

stores. Because of the high demand for and<br />

quick turnover rate of rats, mice, and hamsters,<br />

the majority of pet stores surveyed purchased<br />

their rodents from various vendors rather than<br />

relying on in-house breeding programs. In this<br />

study, 7 of the 12 pet stores purchased animals<br />

from 5 different vendors, while the other 5<br />

stores relied on in-house breeding programs or<br />

various suppliers, either private or commercial<br />

sources. Rodents in those 7 vendor-supplied<br />

stores tested positive for H. nana, while only 3<br />

of the other 5 stores revealed positive rodents<br />

(Table 3).<br />

Evidence for direct transmission of H. nana<br />

as the common route of infection in rodents can<br />

be derived from the concomitant presence of H.<br />

dimimita and H. nana within the same rat cage.<br />

Transmission of H. diminuta requires an arthropod<br />

intermediate host (Roberts and Janovy,<br />

<strong>2000</strong>). Since the same arthropods may serve as<br />

intermediate hosts for both tapeworms, lower<br />

prevalence of H. diminuta and higher prevalence<br />

of H. nana indicate transmission through direct<br />

rather than indirect routes. If H. nana were using<br />

an intermediate host, one would assume the<br />

prevalences of the 2 tapeworms to be nearly<br />

equivalent.<br />

Traditionally, H. nana is considered a tapeworm<br />

of mice (Markell et al., 1999). However,<br />

the public health significance of the higher prevalence<br />

in rats becomes apparent when the type<br />

of pet most often purchased for children is considered.<br />

According to pet store owners, mice are<br />

usually sold as feeder animals, but hamsters and<br />

Table 3. Summary of survey data on Hymenolepis nana in rodents from pet stores in southern Connecticut,<br />

U.S.A.<br />

Pet store Sample<br />

location size<br />

Necropsy animals:]:<br />

Hamden<br />

Wallingford<br />

Meriden<br />

East Haven<br />

North Branford<br />

Fairrield<br />

Orange<br />

Stratford<br />

Naugatuck (a)<br />

Fecal sample||<br />

Naugatuck (b)<br />

Seymore (a)<br />

Seymore (b)<br />

20<br />

17<br />

10<br />

17<br />

15<br />

21<br />

13<br />

18<br />

18<br />

37<br />

23<br />

9<br />

Store prevalence<br />


200 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

rats are more often purchased as pets. Surprisingly,<br />

our results indicate that pet store rats constitute<br />

a more important reservoir for H. nana<br />

than do mice or hamsters. Also, this is the first<br />

report of H. nana from a pet prairie dog, a nontraditional<br />

or exotic animal that is becoming<br />

common in pet stores (Storer and Watson, 1997).<br />

Improper hygiene following handling of all rodents,<br />

including feeders, may lead to transmission.<br />

A pet ownership profile (Teclaw et al.,<br />

1992) showed that approximately 50% of households,<br />

primarily those with children between the<br />

ages of 6-17 years, owned some type of pet.<br />

Further, 2% of Florida AIDS patients interviewed<br />

owned pet rodents, but most health care<br />

providers failed to advise them of possible zoonoses<br />

from their companion animals (Conti et<br />

al., 1995).<br />

Overall, 75% of surveyed pet stores were selling<br />

animals infected with H. nana. Despite the<br />

fact that H. nana infection in rodents is easily<br />

treatable with praziquantel (Harkness and Wagner,<br />

1995), none of the pet stores reported practicing<br />

antihelmintic treatment and control measures.<br />

The combination of high prevalence and<br />

absence of control measures demonstrates that<br />

pet rodents pose a zoonotic threat to pet store<br />

personnel, animal care workers, and customers.<br />

Surveys of human populations, together with<br />

further epidemiological information, are needed<br />

to assess the extent to which this potential health<br />

threat is actually being realized.<br />

Acknowledgments<br />

This work was funded in part by an Interdisciplinary<br />

Research Grant from Quinnipiac <strong>College</strong>.<br />

Robin LePardo assisted in collecting fecal<br />

samples. Kristen E. Richardson, Quinnipiac <strong>College</strong>,<br />

assisted in the preparation of the manuscript.<br />

Literature Cited<br />

Chomel, B. B. 1992. Zoonoses of house pets other<br />

than dogs, cats, and birds. Pediatric Infectious<br />

Disease Journal 11:479-487.<br />

Conti, L., S. Lieb, T. Liberti, M. Wiley-Bayless, K.<br />

Hepburn, and T. Diaz. 1995. Pet ownership<br />

among persons with AIDS in three Florida counties.<br />

American Journal of Public Health 85:1559-<br />

1561.<br />

Cooper, B. T., H. J. Hodgson, and V. S. Chadwick.<br />

1981. Hymenolepiasis: an unusual cause of diarrhea<br />

in Western Europe. Digestion 21:115-116.<br />

Croll, N. A., and T. W. Gyorkos. 1979. Parasitic dis-<br />

Copyright © 2011, The Helminthological Society of Washington<br />

ease in humans: the extent in Canada. Canadian<br />

Medical Association Journal 120:310-312.<br />

Crompton, D. W. T. 1999. How much human helminthiasis<br />

is there in the world? Journal of <strong>Parasitology</strong><br />

85:397-403.<br />

Flores, E. C., S. C. Plumb, and M. C. McNeese.<br />

1983. Intestinal parasitosis in an urban pediatric<br />

clinic population. American Journal of Diseases<br />

of Children 137:754-756.<br />

Fox, J. G., C. E. Newcomer, and H. Ruzmiarek.<br />

1984. Selected zoonoses and other health hazards.<br />

Pages 614-643 in J. G. Fox, B. J. Cohen, and F.<br />

M. Loew, eds. Laboratory Animal Medicine. Academic<br />

Press, Orlando, Florida.<br />

Gerber, C. P., J. B. Rose, and C. N. Haas. 1996.<br />

Sensitive populations: who is at the greatest risk?<br />

International Journal of Food Microbiology 30:<br />

113-123.<br />

Ghandour, A. M., N. Z. Zahid, A. A. Banaja, K. B.<br />

Karnal, and A. I. Bouq. 1995. Zoonotic intestinal<br />

parasites of hamadryas baboons Papio hamadryas<br />

in the western and northern regions of Saudi Arabia.<br />

Journal of Tropical Medicine and Hygiene 98:<br />

431-439.<br />

Guarner, J., T. Matilde-Nava, R. Villasenor-Flores,<br />

and G. Sanchez-Mejorada. 1997. Frequency of<br />

intestinal parasites in adult cancer patients in<br />

Mexico. Archives of Medical Research 28:219-<br />

222.<br />

Harkness, J. E., and J. E. Wagner. 1995. The Biology<br />

and Medicine of Rabbits and Rodents, 4th<br />

ed. Williams & Wilkins, Media, Pennsylvania.<br />

372 pp.<br />

Hendrix, C. M. 1998. Diagnostic Veterinary <strong>Parasitology</strong>,<br />

2nd ed. Mosby, Inc., St. Louis, Missouri.<br />

321 pp.<br />

Hunninen, A. V. 1935. Studies on the life history and<br />

host-parasite relations of Hymenolepis fraterna<br />

(H. nana, van fraterna, Stiles) in white mice.<br />

American Journal of Hygiene 20:414-443.<br />

Jacobs, D. E. 1979. Man and his pets. Pages 174-200<br />

in R. J. Donaldson, ed. Parasites and Western<br />

Man. University Park Press, Baltimore, Maryland.<br />

Jacoby, R. O., and J. G. Fox. 1984. Biology and<br />

diseases of mice. Pages 31-88 in J. G. Fox, B. J.<br />

Cohen, and F. M. Loew, eds. Laboratory Animal<br />

Medicine. Academic Press, Orlando, Florida.<br />

Kappus, K. K., D. D. Juranek, and J. M. Roberts.<br />

1991. Results of testing for intestinal parasites by<br />

state diagnostic laboratories, United <strong>State</strong>s, 1987.<br />

Morbidity and Mortality Weekly Reports Intestinal<br />

Parasite Surveillance Summary 40 (No. SS-<br />

4): 24-45.<br />

Khalil, H. M., S. el Shimi, M. A. Sarwat, A. F.<br />

Fawzy, and A. O. el Sorougy. 1991. Recent<br />

study of Hymenolepis nana infection in Egyptian<br />

children. Journal of the Egyptian Society of Parasitologists<br />

21:293-300.<br />

Little, M. D. 1985. Cestodes (tapeworms). Pages 110-<br />

126 in P. C. Beaver and R. C. Jung, eds. Animal<br />

Agents and Vectors of Human Disease, 5th ed.<br />

Lea and Febiger, Philadelphia, Pennsylvania.<br />

Lussier, G., and F. M. Loew. 1970. Natural Hymenolepis<br />

nana infection in Mongolian gerbils (Mer-


tones ungiiiculatus). Canadian Veterinary Journal<br />

11:105-107.<br />

Markell, E. K., D. T. John, and W. A. Krotoski.<br />

1999. Markell and Voge's Medical <strong>Parasitology</strong>,<br />

8th ed. W. B. Saunders Co., Philadelphia, Pennsylvania.<br />

501 pp.<br />

Meloni, B. P., R. C. A. Thompson, R. M. Hopkins,<br />

J. A. Reynoldson, and M. Gracey. 1993. The<br />

prevalence of Giardia and other intestinal parasites<br />

in children, dogs and cats from Aboriginal<br />

communities in the Kimberly. Medical Journal of<br />

Australia 158:157-159.<br />

Neva, F. A., and H. W. Brown. 1994. Basic Clinical<br />

<strong>Parasitology</strong>, 6th ed. Appleton and Lange, Norwalk,<br />

Connecticut. 356 pp.<br />

Roberts, L. S., and J. Janovy, Jr. <strong>2000</strong>. Gerald D.<br />

Schmidt and Larry S. Roberts' Foundations of<br />

<strong>Parasitology</strong>, 6th ed. McGraw-Hill, Boston, Massachusetts.<br />

<strong>67</strong>0 pp.<br />

Sahin, I. 1979. Parasitosis and zoonosis in mice and<br />

rats caught in and around Beytepe Village. Mikrobiyoloji<br />

Bulteni 13:283-290.<br />

Seaton, J. R. 1979. Cestodes and trematodes. Pages<br />

114-132 in R. J. Donaldson, ed. Parasites and<br />

DUCLOS AND RICHARDSON—HYMENOLEPIS NANA IN PETS 201<br />

NEW BOOK AVAILABLE<br />

Western Man. University Park Press, Baltimore,<br />

Maryland.<br />

Stone, W. B., and R. D. Manwell. 1966. Potential<br />

helminth infections in humans from pet or laboratory<br />

mice and hamsters. Public Health Reports<br />

31:647-653.<br />

Storer, P., and L. Watson. 1997. Prairie Dog Primer.<br />

Country Storer Enterprises, Columbus, Texas. 76<br />

pp.<br />

Teclaw, R., J. M. Mendlein, P. Garbe, and P. Mariolis.<br />

1992. Characteristics of pet populations and<br />

households in the Purdue <strong>Comparative</strong> Oncology<br />

Program catchment area, 1988. Journal of the<br />

American Veterinary Medical Association 210:<br />

1725-1729.<br />

Turner, J. A. 1975. Other cestode infections. Pages<br />

708-744 in J. A. Hubbert, W. F. McCulloch, and<br />

P. R. Schnurrenberger, eds. Diseases Transmitted<br />

from Animals to Man, 6th ed. Charles C. Thomas,<br />

Springfield, Illinois.<br />

Wilairatana, P., B. Radomyos, R. Phraevanich, W.<br />

Plooksawasdi, P. Chanthavanich, C. Viravan,<br />

and S. Looareesuwan. 1996. Intestinal sarcocystosis<br />

in Thai laborers. Southeast Asian Journal of<br />

Tropical Medicine and Public Health 27:43-46.<br />

Echinostomes as Experimental Models for Biological Research. Edited by Bernard Fried and<br />

Thaddeus K. Graczyk. <strong>2000</strong>. Kluwer Academic Publications, Dordrecht, The Netherlands. 284 pp.<br />

ISBN 0-7923-6156-3. Hardcover. US$ 150.00/NLG 250.00/GBP 88.00. Abstract: ". . . Considerable<br />

but scattered literature has been published on the subject of echinostomes and a synthesis of<br />

this wide range of topics has now been achieved with the publication of this book, which represents<br />

a wide range of topics in experimental biology related to the use of echinostomes as laboratory<br />

models. It will have a special appeal to advanced undergraduates and graduate students in parasitology<br />

and should also appeal to professional parasitologists, physicians, veterinarians, wildlife<br />

disease biologists, and any biomedical scientists interested in new model systems for studies in<br />

experimental biology."<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 202-209<br />

Seasonal Occurrence and Community Structure of Helminth<br />

Parasites from the Eastern American Toad, Bufo americanus<br />

americanus., from Southeastern Wisconsin, U.S.A.<br />

MATTHEW G. BOLEK' AND JAMES R. COGGINS<br />

Department of Biological Sciences, University of Wisconsin-Milwaukee, Milwaukee, Wisconsin, 53201,<br />

U.S.A. (e-mail: coggins@csd.uwm.edu)<br />

ABSTRACT: From April to September 1996, 47 American toads, Bufo americanus americanus Holbrook, were<br />

collected from Waukesha County, Wisconsin, U.S.A., and examined for helminth parasites. Forty-six (98%) of<br />

47 toads were infected with 1 or more helminth species. The component community consisted of 6 species, 3<br />

direct-life-cycle nematodes, and 3 indirect-life-cycle helminths (2 trematodes and 1 metacestode). Totals of 2,423<br />

individual nematodes (92%), 45 trematodes (2%), and 155 cestodes (6%) were found, with infracommunities<br />

being dominated by skin-penetrating nematodes. A significant correlation existed between wet weight and overall<br />

helminth abundance, excluding larval platyhelminths. Helminth populations and communities were seasonally<br />

variable but did not show significant differences during the year. However, a number of species showed seasonal<br />

variations in location in the host, and these variations were related to recruitment period.<br />

KEY WORDS: Bufo americanus, American toad, Cosmocercoides variahilis, Rhabdias americanus, Oswaldocruzia<br />

pipiens, Mesocestoidex sp., Gorgoderina sp., Trematoda, Nematoda, Cestoda, echinostome metacercariae,<br />

seasonal study, Wisconsin, U.S.A.<br />

American toads, Bufo americanus americanus<br />

Holbrook, 1836, are large, thick-bodied terrestrial<br />

anurans found in North America near<br />

marshes, oak savannas, semiopen coniferous and<br />

deciduous forests, and agricultural areas. They<br />

range from Labrador and Hudson Bay to eastern<br />

Manitoba, south to eastern Oklahoma and the<br />

coastal plains, and are distributed throughout<br />

Wisconsin (Vogt, 1981). Toads are active foragers,<br />

differing from most anurans that are sitand-wait<br />

predators (Scale, 1987). Although a<br />

number of surveys and natural history studies<br />

on the helminths and ecology of toads exists<br />

(Bouchard, 1951; Odlaug, 1954; Ulmer, 1970;<br />

Ulmer and James, 1976; Williams and Taft,<br />

1980; Coggins and Sajdak, 1982; Joy and Bunten,<br />

1997), no studies have used measures of<br />

helminth communities. Here we report on the<br />

helminth infracommunity and component community<br />

structure in American toads from southeastern<br />

Wisconsin.<br />

Materials and Methods<br />

American toads were collected from April to November<br />

of 1996 by driving 4.0-km sections of highways<br />

N and <strong>67</strong> (42°54'N, 88°29'W) in Eagle, Wau-<br />

1 Corresponding author. Current address: Department<br />

of Veterinary Pathobiology, Purdue University,<br />

West Lafayette, Indiana 47907, U.S.A. (e-mail:<br />

mgb @ vet.purdue.edu).<br />

202<br />

Copyright © 2011, The Helminthological Society of Washington<br />

kesha County, Wisconsin, U.S.A., during the night and<br />

collecting individuals as they crossed roads. Animals<br />

were placed in plastic containers, transported to the<br />

laboratory, stored at 4°C, and killed in MS-222 (ethyl<br />

m-aminobenzoate methane sulfonic acid) within 72 hr<br />

of capture. Snout-vent length (SVL) and wet weight<br />

(WW) were recorded for each individual. Toads were<br />

individually toe-clipped and frozen. At necropsy, the<br />

digestive tracts, limbs, body wall musculature, and internal<br />

organs were examined for helminth parasites.<br />

Each organ was individually placed in a Petri dish and<br />

examined under a stereomicroscope. The body cavity<br />

was rinsed with distilled water into a Petri dish and<br />

the contents examined. All individuals were sexed by<br />

gonad inspection during necropsy. Worms were removed<br />

and fixed in alcohol-formaldehyde-acetic acid<br />

or formalin. Trematodes and cestodes were stained<br />

with acetocarmine, dehydrated in a graded ethanol series,<br />

cleared in xylene, and mounted in Canada balsam.<br />

Nematodes were dehydrated to 70% ethanol, cleared<br />

in glycerol, and identified as temporary mounts. Echinostome<br />

metacercariae were badly damaged during<br />

necropsy, and these were identified but not retained.<br />

Prevalence, mean intensity, and abundance are according<br />

to Bush et al. (1997). Mean helminth species richness<br />

is the sum of helminth species per individual amphibian,<br />

including noninfected individuals, divided by<br />

the total sample size. All values are reported as the<br />

mean ± 1 SD. Undigested stomach contents were<br />

identified to class or order following Borror et al.<br />

(1989). Stomach contents are reported as a percent =<br />

the number of arthropods in a given class or order,<br />

divided by the total number of arthropods recovered<br />

X 100. Voucher specimens have been deposited in the<br />

helminth collection of the H. W. Manter Laboratory<br />

(HWML), University of Nebraska <strong>State</strong> Museum, Lincoln,<br />

Nebraska, U.S.A. (accession numbers HWML


BOLEK AND COGGINS—HELMINTH COMMUNITIES IN TOADS 203<br />

Table 1. Prevalence, mean intensity, mean abundance, and total numbers of helminths found in 47 specimens<br />

of Bufo americanus americanus.<br />

Species<br />

Trematoda<br />

Echinostome metacercariae*<br />

Gorgoderina sp.<br />

Cestoda<br />

Mesocestoides sp.*<br />

Nematoda<br />

Oswaldocruzia pipiens<br />

Cosmocercoides variabilis<br />

Rhahdias americanus<br />

Underestimate.<br />

Prevalence, Mean intensity<br />

No. (%) ± 1 SD (range)<br />

3 (6.3)<br />

1 (2.1)<br />

13 ± 19 (1-35)<br />

6 ± 0 (6)<br />

6 (12.7) 25.8 ± 22 (12-70)<br />

15051, male Oswaldocruzia pipiens; 15052, male Cosmocercoides<br />

variabilis; 15053, Rhabdias americanus;<br />

15054, Gorgoderina sp.; 15055, Mesocestoides sp.).<br />

The chi-square test for independence was calculated<br />

to compare differences in prevalence among host sex,<br />

seasonal differences in prevalence, and seasonal differences<br />

in location of nematodes in the host. Yates'<br />

adjustment for continuity was used when sample sizes<br />

were low, and a single-factor, independent-measures<br />

analysis of variance was used to compare among seasonal<br />

differences in mean intensity and mean helminth<br />

species richness (Sokal and Rohlf, 1981). Student's ttest<br />

was used to compare differences in mean intensity<br />

and mean helminth species richness between sex of<br />

hosts. Approximate f-tests were calculated when variances<br />

were heteroscedastic (Sokal and Rohlf, 1981).<br />

Pearson's correlation was used to determine relationships<br />

among host SVL and WW and abundance of<br />

helminth parasites, excluding larval platyhelminths.<br />

Pearson's con-elation was calculated for host SVL and<br />

WW and helminth species richness per individual amphibian.<br />

Because WW gave a stronger correlation than<br />

SVL in each case, it is the only parameter reported.<br />

Because of low sample sizes during certain collection<br />

periods, data were pooled on a bimonthly basis to form<br />

samples of 15 to 16 toads per season. Larval platyhelminths<br />

were not included in the seasonal analysis,<br />

because they can accumulate throughout an amphibian's<br />

life.<br />

Results<br />

A total of 47 American toads, 28 males and<br />

19 females, was collected. The overall mean<br />

SVL and WW of toads was 56.6 ± 12.5 mm<br />

(range = 26.2-72.6 mm) and 26.6 ± 13.5 g<br />

(range = 2.19-55.5 g), respectively. No significant<br />

difference existed in numbers of male<br />

toads and female toads collected throughout the<br />

No. of<br />

Mean worms<br />

abundance recov-<br />

± 1 SD ered Location<br />

0.8 ± 5.1<br />

0.1 ± 0.9<br />

41 (87) 8.5 ± 7 (1-31) 7.4 ± 7.1<br />

43(91) 32.3 ± 31.5 (1-135) 29.6 ±31.5<br />

39 Kidneys, body cavity<br />

6 Bladder<br />

155 Leg muscles<br />

349 Small intestine<br />

1,392 Lungs, body cavity,<br />

large and small intestine<br />

43(91) 15.8 ± 17.9 (1-75) 14.5 ± 17.7 682 Lungs, body cavity<br />

year (x2 = 1.72, P > 0.05). Although female<br />

toads were larger (58.2 ± 15.4 mm) and heavier<br />

(30.5 ± 17.3 g) than males (55.5 ± 10.2 mm,<br />

23.9 ± 9.7 g), these differences were not significant<br />

(t = 0.73, P > 0.05, t's = 1.50, P > 0.05).<br />

Stomach contents analyses revealed that the<br />

toads fed mostly on ants (98%), with beetles and<br />

other terrestrial arthropods representing a small<br />

portion of the diet (2%).<br />

Forty-six (98%) of 47 toads were infected<br />

with 1 or more species of helminths. The component<br />

community consisted of 6 species, 3 direct-life-cycle<br />

nematodes, and 3 indirect-life-cycle<br />

helminths (2 trematodes and 1 metacestode).<br />

Overall mean helminth abundance, excluding<br />

larval platyhelminths, was 55.7 ± 45.3 worms<br />

per toad infracommunity (range = 0-180). Prevalence<br />

was highest for nematodes, ranging from<br />

91% for Cosmocercoides variabilis Harwood,<br />

1930, and Rhabdias americanus Baker, 1978, to<br />

87% for Oswaldocruzia pipiens Walton, 1929.<br />

Prevalence for indirect life cycle parasites was<br />

generally low, being highest for the cestode Mesocestoides<br />

sp. (12.7%) and lowest for Gorgoderina<br />

sp. (2.1%) (Table 1). No significant differences<br />

existed in prevalence between male and<br />

female toads for any of the 6 helminth species<br />

recovered. Mean intensity differed significantly<br />

only in Mesocestoides sp., being higher in male<br />

(32.7 ± 32) than in female toads (19 ± 4.6, /;<br />

= 4.70, P < 0.05).<br />

Mean helminth species richness was 2.9 ± 0.9<br />

Copyright © 2011, The Helminthological Society of Washington


20-1 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

.1 u<br />

ft,<br />

cc<br />

"e<br />

1<br />

"S<br />

ffi<br />

o<br />

L.<br />

a><br />

ja<br />

S<br />

3<br />

O -<br />

5 -<br />

4 -<br />

Figure 1.<br />

3 - -*• • 4*MHN»*++ •» * ** *<br />

2 -<br />

** «•* •<br />

1 - •*• •*•<br />

n .<br />

*<br />

* ++ * •<br />

, A .4,.- 4 1<br />

20 40<br />

Wet Weight in Grams<br />

Figure 3.<br />

S Body Cavity B Lungs<br />

100 -r D Small Intestine • Large Intestine<br />

= 405<br />

Aug-Sep<br />

60<br />

0<br />

Number of ¥<br />

1.80-1<br />

160-<br />

140-<br />

120-<br />

100 -<br />

80-<br />

60-<br />

40-<br />

Figure 2. n<br />

• Males<br />

D Females + D<br />

• D<br />

& U<br />

•<br />

- * DV**<br />

^ •<br />

20-<br />

"P<br />

A . ctk<br />

^ •* tl *<br />

* * n i<br />

20 40<br />

Wet Weight in Grams<br />

w<br />

1<br />

100-r<br />

Figure 4.<br />

I Body Cavity D Lungs<br />

60<br />

= 261<br />

Figures 1—4. 1. Wet weight versus number of helminth species per individual of the American toad,<br />

Bufo americanus americanus, r = 0.31, P < 0.05. 2. Wet weight versus total helminth abundance, excluding<br />

larval platyhelminths, in American toads, B. a. americanus, overall r = 0.47, P < 0.01; female toads r =<br />

0.57, P < 0.01; and male toads r = 0.20, P > 0.05. 3. Seasonal distribution of the relative proportions of<br />

Cosmocercoides variabilis recovered in the body cavity, lungs, small intestine, and large intestine of B. a.<br />

americanus. N equals number of nematodes recovered in each sampling period. 4. Seasonal distribution<br />

of the relative proportions of Rhabdias americanus recovered in the body cavity and lungs of B. a. americanus.<br />

N equals number of nematodes recovered in each sampling period.<br />

species per toad. Infections with multiple species<br />

were common, with 0, 1,2, 3, 4, and 5 species<br />

occurring in 1, 2, 6, 30, 7, and 1 host, respectively.<br />

No statistically significant differences in<br />

mean helminth species richness were found between<br />

male (3.0 ± 0.86) and female toads (2.8<br />

Copyright © 2011, The Helminthological Society of Washington<br />

± 0.85, t = 0.83, P > 0.05). There was a significant<br />

positive correlation between WW and<br />

helminth species richness per toad (r = 0.31, P<br />

< 0.05, Fig. 1). However, this relationship became<br />

insignificant when a single uninfected toad<br />

was removed (r = 0.20, P > 0.05). A significant


BOLEK AND COGGINS—HELMINTH COMMUNITIES IN TOADS 205<br />

Table 2. Seasonal prevalence and mean intensity of 3 species of nematodes in Bufo americanus americanus.<br />

Species Apr Jun-Jul Aug-Sep Statistic<br />

Cosmocercoides variahilis<br />

Oswaldocruzia pipiens<br />

Rhabdias americanus<br />

Prevalence<br />

Mean intensity ±<br />

Prevalence<br />

Mean intensity ±<br />

Prevalence<br />

Mean intensity ±<br />

1 SD<br />

1 SD<br />

1 SD<br />

93%<br />

30.4<br />

93%<br />

7<br />

93%<br />

17.8<br />

positive correlation existed between WW and<br />

overall helminth abundance, excluding larval<br />

platyhelminths (r = 0.47, P < 0.01, Fig. 2), although<br />

this correlation was only significant for<br />

female toads (r = 0.57, P < 0.01) and not for<br />

males (r = 0.20, P > 0.05). Similar results were<br />

obtained for C. vanabilis (r = 0.41, P < 0.01)<br />

and O. pipiens (r = 0.43, P < 0.01), whereas<br />

no significant correlation was observed for R.<br />

americanus (r = 0.27, P > 0.05). When these<br />

analyses were performed separately for male and<br />

female hosts, significant positive correlations occurred<br />

only in female toads (C. variabilis, r —<br />

0.52, P < 0.05; O. pipiens, r = 0.58, P < 0.01;<br />

R. americanus, r = 0.54, P < 0.05).<br />

Only 1 male toad was infected with 6 Gorgoderina<br />

sp. during the early spring (April) collection.<br />

There was no significant seasonal difference<br />

in prevalence or mean intensity for any<br />

140<br />

Number of Osn>aldocruzia pipiens<br />

Figure 5. Number of Oswaldocruzia pipiens versus<br />

Cosmocercoides variabilis in Bufo americanus<br />

americanus, overall r = 0.54, P < 0.01; female toads<br />

r = 0.58, P < 0.01; and male toads r = 0.51, P <<br />

0.01.<br />

(14/15)<br />

± 28.5<br />

(14/15)<br />

± 6.7<br />

(14/15)<br />

± 21.6<br />

88% (14/16)<br />

40 ± 31.7<br />

88%<br />

10<br />

100%<br />

10.7<br />

(14/16)<br />

± 8.1<br />

(16/16)<br />

±11.2<br />

94%<br />

27<br />

81%<br />

8.3<br />

81%<br />

20<br />

(15/16)<br />

± 34.6<br />

(13/16)<br />

± 6.1<br />

(13/16)<br />

± 19.8<br />

X2<br />

F<br />

X2<br />

F<br />

X2<br />

F<br />

= 0.5<br />

= 0.65<br />

= 1.1<br />

= 0.63<br />

= 3.7<br />

= 1.12<br />

P > 0.05<br />

P > 0.05<br />

P > 0.05<br />

P > 0.05<br />

P > 0.05<br />

P > 0.05<br />

of the nematodes recovered (Table 2). Prevalence<br />

was highest during early spring for O. pipiens<br />

(93%), early summer (June-July) for R.<br />

americanus (100%), and late summer-early fall<br />

(August-September) for C. vanabilis (94%).<br />

Mean intensities were higher during early summer<br />

for C. variabilis and O. pipiens and in late<br />

summer-early fall for R. americanus. Seasonally,<br />

there was a significant difference in location<br />

of C. variabilis (x2 = 556, P < 0.01), and R.<br />

americanus (x2 = 232, P < 0.01). Cosmocercoides<br />

variabilis occurred in greater numbers in<br />

the body cavity and lungs during the early spring<br />

and in the small and large intestines during early<br />

summer and late summer-early fall collections<br />

(Fig. 3). Individuals of Rhabdias americanus<br />

were found primarily in the lungs during early<br />

spring and progressively increased in numbers<br />

in the body cavity during early summer and late<br />

summer-early fall collections (Fig. 4). The nematode<br />

O. pipiens showed no seasonal variation<br />

in location and was found in the small intestine<br />

throughout the year. Oswaldocruzia pipiens was<br />

significantly correlated with C. variabilis (r =<br />

0.54, P < 0.01) in both male (r = 0.51, P <<br />

0.01) and female toads (r = 0.58, P < 0.01)<br />

(Fig. 5) but was not significantly correlated with<br />

R. americanus (r = 0.23, P > 0.05) in either<br />

male (r = 0.23, P > 0.05) or female toads (r =<br />

0.24, P > 0.05).<br />

Mean species richness varied throughout the<br />

year, being highest (3.26 ± 0.79) in early spring,<br />

intermediate (2.88 ± 0.71) in early summer, and<br />

lowest during the late summer-early fall collection<br />

(2.62 ± 0.95), although these differences<br />

were not statistically significant (F = 2.3, P ><br />

0.05). Wet weight of toads showed similar seasonal<br />

variability, but these differences were also<br />

not significant (F = 2.99, P > 0.05).<br />

Discussion<br />

The component community of Wisconsin<br />

toads was similar to those of other published re-<br />

Copyright © 2011, The Helminthological Society of Washington


206 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

ports (Bouchard, 1951; Odlaug, 1954; Ulmer,<br />

1970; Ulmer and James, 1976; Williams and<br />

Taft, 1980; Coggins and Sajdak, 1982; Joy and<br />

Bunten, 1997; Yoder, 1998). Toad helminth infracommunities<br />

were dominated by 3 species of<br />

skin-penetrating nematodes with high overall<br />

prevalence and mean intensities and with few<br />

toads infected by indirect-life-cycle parasites.<br />

Bladder flukes of the genus Gorgoderina are<br />

common parasites of amphibians, but few lifecycle<br />

studies exist (Prudhoe and Bray, 1982).<br />

The life cycles of Gorgoderina attenuata Stafford,<br />

1902, and Gorgoderina vitelliloba Olsson,<br />

1876, have been determined. For these species,<br />

amphibians acquire infection by feeding on<br />

semiaquatic insect larvae or tadpoles, and the<br />

worms excyst in the stomach and migrate to the<br />

kidneys and bladder (Rankin, 1939; Smyth and<br />

Smyth, 1980). The low prevalence (2.1%) and<br />

intensity (6) of Gorgoderina sp. observed in our<br />

study were not surprising, since analyses of<br />

stomach contents revealed few kinds of arthropods.<br />

This is characteristic of actively foraging<br />

species such as toads. Ants made up the largest<br />

portion of the diet, with beetles and other terrestrial<br />

arthropods being found less frequently.<br />

Kirkland (1904) also found that ants and beetles<br />

made up the greatest portion of the diet of 149<br />

toads from New England, U.S.A. These results<br />

are similar to other investigations on diet of species<br />

ofBufo (Toft, 1981; Collins, 1993; Indraneil<br />

and Martin, 1998), where ants and beetles appeared<br />

to be an important food item in the diet<br />

of toads. Toft (1981) reported that ants made up<br />

64% to 91% of the arthropods consumed by 3<br />

South American toad species, while Collins<br />

(1993) mentioned that ants and beetles were important<br />

items in the diet of 5 species of Bufo<br />

from Kansas, U.S.A. Because most trematodes<br />

of amphibians use aquatic or semiaquatic arthropods<br />

as intermediate hosts (Prudhoe and Bray,<br />

1982), these observations may indicate why<br />

toads usually have low species richness and<br />

prevalence of adult trematodes (Williams and<br />

Taft, 1980; Coggins and Sajdak, 1982; McAllister<br />

et al., 1989; Goldberg and Bursey, 1991a,<br />

1991b, 1996; Goldberg et al., 1995; Bursey and<br />

Goldberg, 1998).<br />

The most commonly occurring nematode was<br />

C. variabilis, with a total of 1,392 worms recovered.<br />

Vanderburgh and Anderson (1987)<br />

studied the seasonal transmission of this species<br />

in American toads from Ontario, Canada. They<br />

Copyright © 2011, The Helminthological Society of Washington<br />

observed J4 larvae in the lungs during the breeding<br />

season and adult worms in the rectum of<br />

toads throughout the year. They suggested that<br />

toads may acquire C. variabilis soon after<br />

emerging in the spring and that transmission<br />

may decline during summer and fall. However,<br />

they stated that this may have been an artifact<br />

of sampling, because all toads collected after the<br />

breeding season were from another location and<br />

may have had a lower prevalence and mean intensity<br />

of C. variabilis. Vanderburgh and Anderson<br />

(1987) also observed larvae in the lungs<br />

of 5 toads collected in October of the following<br />

year and concluded that transmission probably<br />

occurs throughout the year.<br />

Data from the present study suggest that the<br />

breeding period may be important in transmission<br />

of C. variabilis in adult toads. All our toads<br />

were collected from the same general location<br />

and had high prevalence and mean intensities<br />

throughout the year. Although the differences<br />

were not significant, mean intensity increased<br />

after the breeding season and decreased during<br />

the late summer-early fall collection. Ten percent<br />

of the worms recovered during April were<br />

located in the body cavity, 37% were located in<br />

the lungs, and 28% and 24% were located in the<br />

small and large intestine, respectively. Subsequent<br />

sampling revealed that only 1 toad collected<br />

during June had 6 larvae in the lungs,<br />

with all other worms being recovered from the<br />

small and large intestine. Baker's (1978a) study<br />

on the life cycle of Cosmocercoides dukae Holl,<br />

1928 (=C. variabilis} in toads revealed that 8-<br />

10 days are required for larvae to reach the<br />

lungs, and more than 30 days at 14-18°C to migrate<br />

to the rectum and develop to a gravid<br />

stage. Our observations suggest that toads became<br />

infected during the breeding season, and<br />

there appeared to be a decline during summer<br />

and early fall. Unfortunately, no toads were collected<br />

during October, and therefore it is not<br />

known if infection may occur during the fall (but<br />

see below). All adult female worms recovered<br />

throughout the year were gravid, indicating that<br />

eggs were being produced from April through<br />

September.<br />

The second most frequently recovered nematode<br />

was R. americanus, primarily a parasite of<br />

toads (Baker, 1979a, 1987). Few studies exist on<br />

the seasonal occurrence of species of Rhabdias<br />

in amphibians (Lees, 1962; Plasota, 1969; Baker,<br />

1979b). Lees (1962) studied the seasonal occur-


ence of Rhabdias bufonis Schrank, 1788, in its<br />

host, Rana temporaria Linnaeus, 1758, in England,<br />

and Baker (1979b) studied the seasonal occurrence<br />

of Rhabdias ranae Walton, 1929, in the<br />

wood frog, Rana sylvatica Le Conte, 1825, in<br />

Canada. Prevalence and intensities in these species<br />

were lowest during summer and highest in<br />

spring and early fall. Baker (1979b) observed<br />

many subadult worms in the body cavity of<br />

wood frogs during late summer and early fall,<br />

with no worms being found in the body cavity<br />

during early spring and few worms in the fall.<br />

Worms occurred in the lungs during early spring<br />

and fall, while few were found in the lungs during<br />

late summer and early fall. Baker (1979b)<br />

concluded that transmission of R. ranae occurs<br />

throughout the summer and early fall, with<br />

worms maturing in the lungs during the fall and<br />

overwintering in their hosts.<br />

During this study, no significant differences in<br />

prevalence or mean intensity were observed<br />

throughout the collection period, although the<br />

location (lungs or body cavity) of worms varied<br />

during the year. Most R. americanus recovered<br />

in April occurred in the lungs, with numbers of<br />

worms in the lungs decreasing during early summer<br />

(June-July) and late summer-early fall (August-September).<br />

In contrast, the number of<br />

worms in the body cavity increased during the<br />

June-July and August-September collections.<br />

Therefore, transmission of this species occurred<br />

during the summer and early fall, with worms<br />

overwintering in their host. These results are<br />

consistent with earlier studies of seasonal distribution<br />

of other species of Rhabdias (Lees, 1962;<br />

Baker, 1979b).<br />

Oswaldocruzia pipiens also had high prevalence<br />

but lower mean intensities than the other<br />

2 species of nematodes recovered. Because of<br />

its fast migration, reaching the stomach and<br />

small intestine within 1 to 3 days of infection<br />

(Baker, 1978b), differences in location of these<br />

worms within the host were not observed. Prevalence<br />

and mean intensity were variable but not<br />

significant over the course of this study. Baker<br />

(1978b), in a seasonal study of O. pipiens in<br />

wood frogs, observed peak prevalence and intensity<br />

during spring (May-June) and early fall<br />

(September-October) in Ontario, Canada. He<br />

stated that worms overwintered in the host and<br />

that transmission occurred in early spring, with<br />

an initial decline during early summer and continued<br />

transmission during summer and early<br />

BOLEK AND COGGINS—HELMINTH COMMUNITIES IN TOADS 207<br />

fall. A significant positive correlation existed for<br />

this species and C. variabilis but not for R.<br />

americanus, which suggested that toads became<br />

infected with O. pipiens and C. variabilis during<br />

the same time and in the same places, while infection<br />

with R. americanus occurred at a later<br />

time during the summer. Hosts that contain different<br />

species of adult parasites that co-occur in<br />

an infracommunity may contaminate an area<br />

when they release eggs in their hosts' feces.<br />

Therefore, acquisition of a certain parasite species<br />

by a host may often be coupled with the<br />

acquisition of other species into the infracommunity.<br />

The spring breeding period may also be<br />

important in the recruitment of O. pipiens in<br />

toads at our study site. Because toads were not<br />

collected during October, it is not known if recruitment<br />

occurs during this time. If infection<br />

occurs during the fall, as Baker's (1978b) data<br />

for wood frogs imply, C. variabilis may also be<br />

acquired in the fall.<br />

Significant, positive relationships between<br />

WW and species richness and abundance were<br />

observed in toad helminth communities. However,<br />

significant relationships between WW and<br />

abundance were observed only in female toads.<br />

Interestingly, the largest toads collected were females,<br />

and these had the highest intensities of<br />

helminths; therefore, they may provide a greater<br />

surface area for colonization by skin-penetrating<br />

nematodes. Similar observations were reported<br />

by Me Alpine (1997) for female leopard frogs,<br />

which were significantly larger than males. Although<br />

species richness also showed a significant<br />

positive relationship with WW, once the<br />

single noninfected individual was excluded from<br />

the calculation, the relation became nonsignificant.<br />

Therefore, no conclusions can be drawn<br />

from this relationship. Seasonal variance in species<br />

richness was not significant in B. a. americanus,<br />

with toad infracommunities being dominated<br />

by 3 skin-penetrating nematodes throughout<br />

the year. The toad's terrestrial habitat and<br />

diet of ants and beetles may be important in excluding<br />

transmission of adult and larval trematodes,<br />

unlike other anurans such as semiaquatic<br />

species of Rana, which spend more time in an<br />

aquatic environment and have a broader diet of<br />

semiaquatic invertebrates (Muzzall, 1991;<br />

McAlpine, 1997; Bolek, 1998).<br />

Acknowledgments<br />

We thank Melissa Ewert and Luke Bolek for<br />

help in collecting toads. We also thank 2 anon-<br />

Copyright © 2011, The Helminthological Society of Washington


208 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

ymous reviewers and the editors, Drs. W. A.<br />

Reid and J. W. Reid, for improvements on an<br />

earlier draft of the manuscript.<br />

Literature Cited<br />

Baker, M. R. 1978a. Transmission of Cosmocercoides<br />

dukae (Nematoda: Cosmocercoidea) to amphibians.<br />

Journal of <strong>Parasitology</strong> 64:765-766.<br />

. 1978b. Development and transmission of Oswaldocruzia<br />

pipiens Walton, 1929 (Nematoda:<br />

Trichostrongylidae) in amphibians. Canadian<br />

Journal of Zoology 56:1026-1031.<br />

. 1979a. The-free living and parasitic development<br />

of Rhabdias spp. (Nematoda: Rhabdiasidae)<br />

in amphibians. Canadian Journal of Zoology 57:<br />

161-178.<br />

. 1979b. Seasonal population changes in Rhabdias<br />

ranae Walton, 1929 (Nematoda: Rhabdiasidae)<br />

in Rana sylvatica of Ontario. Canadian Journal<br />

of Zoology 57:179-183.<br />

-. 1987. Synopsis of the Nematoda Parasitic in<br />

Amphibians and Reptiles. Occasional Papers in<br />

Biology, Memorial University of Newfoundland,<br />

St. John's, Newfoundland, Canada. 325 pp.<br />

Bolek, M. G. 1998. A seasonal and comparative study<br />

of helminth parasites in nine Wisconsin amphibians.<br />

M.S. Thesis, University of Wisconsin-Milwaukee.<br />

134 pp.<br />

Borror, D. J., C. A. Triplehorn, and N. F. Johnson.<br />

1989. An Introduction to the Study of Insects, 6th<br />

ed. Harcourt Brace <strong>College</strong> Publishers, Philadelphia,<br />

Pennsylvania. 875 pp.<br />

Bouchard, J. L. 1951. The platyhelminths parasitizing<br />

some northern Maine Amphibia. Transactions of<br />

the American Microscopical Society 70:245-250.<br />

Bursey, C. R., and S. R. Goldberg. 1998. Helminths<br />

of the Canadian toad, Bufo hemiophrys (Amphibia:<br />

Anura), from Alberta, Canada. Journal of <strong>Parasitology</strong><br />

84:617-618.<br />

Bush, A. O., K. D. Lafferty, J. M. Lotz, and A. W.<br />

Shostak. 1997. <strong>Parasitology</strong> meets ecology on its<br />

own terms: Margolis et al. revisited. Journal of<br />

<strong>Parasitology</strong> 83:575-583.<br />

Coggins, J. R., and R. A. Sajdak. 1982. A survey of<br />

helminth parasites in the salamanders and certain<br />

anurans from Wisconsin. Proceedings of the Helminthological<br />

Society of Washington 49:99-102.<br />

Collins, J. T. 1993. Amphibians and Reptiles in Kansas,<br />

3rd ed. The University of Kansas Museum of<br />

Natural History, Lawrence, Kansas, Public Education<br />

Series 13. 397 pp.<br />

Goldberg, S. R., and C. R. Bursey. 199la. Helminths<br />

of the red-spotted toad, Bufo punctatus (Anura:<br />

Bufonidae), from southern Arizona. Journal of the<br />

Helminthological Society of Washington 58:261-<br />

269.<br />

, and . 199 Ib. Helminths of three toads,<br />

Bufo alvarius, Bufo cognatus (Bufonidae), and<br />

Scaphiopus couchii (Pelobatidae), from southern<br />

Arizona. Journal of the Helminthological Society<br />

of Washington 58:142-146.<br />

, and . 1996. Helminths of the oak toad<br />

Copyright © 2011, The Helminthological Society of Washington<br />

(Bufo quercicus, Bufonidae) from Florida<br />

(U.S.A.). Alytes 14:122-126.<br />

-, and I. Ramos. 1995. The component<br />

parasite community of three sympatric toad species,<br />

Bufo cognatus, Bufo debilis (Bufonidae), and<br />

Spea multiplicata (Pelobatidae) from New Mexico.<br />

Journal of the Helminthological Society of<br />

Washington 62:57-61.<br />

Indraneil, D., and G. N. Martin. 1998. Bufo fergusonii.<br />

Ecology. Herpetological'Review 29:164.<br />

Joy, J. E., and C. A. Bunten. 1997. Cosmocercoides<br />

variabilis (Nematoda: Cosmocercoidea) populations<br />

in the eastern American toad, Bufo a. americanus<br />

(Salienta: Bufonidae), from western West<br />

Virginia. Journal of the Helminthological Society<br />

of Washington 64:102-105.<br />

Kirkland, A. H. 1904. Usefulness of the American<br />

toad. Farm Bulletin, U.S. Department of Agriculture,<br />

Washington, D.C., 196:1-16.<br />

Lees, E. 1962. The incidence of helminth parasites in<br />

a particular frog population. <strong>Parasitology</strong> 52:95—<br />

102.<br />

McAllister, C. T., S. J. Upton, and D. B. Conn.<br />

1989. A comparative study of endoparasites in<br />

three species of sympatric Bufo (Anura: Bufonidae),<br />

from Texas. Proceedings of the Helminthological<br />

Society of Washington 56:162-1<strong>67</strong>.<br />

McAlpine, D. F. 1997. Helminth communities in bullfrogs<br />

(Rana catesbeiana), green frogs (Rana<br />

clamitans), and leopard frogs (Rana pipiens) from<br />

New Brunswick, Canada. Canadian Journal of Zoology<br />

75:1883-1890.<br />

Muzzall, P. M. 1991. Helminth infracommunities of<br />

the frogs Rana catesbeiana and Rana clamitans<br />

from Turkey Marsh, Michigan. Journal of <strong>Parasitology</strong><br />

77:366-371.<br />

Odlaug, T. O. 1954. Parasites of some Ohio Amphibia.<br />

Ohio Journal of Science 54:126-128.<br />

Plasota, K. 1969. The effect of some ecological factors<br />

on the parasitofauna of frogs. Acta Parasitologica<br />

Polonica 16:47-60.<br />

Prudhoe, O. B. E., and R. A. Bray. 1982. Platyhelminth<br />

Parasites of the Amphibia. British Museum<br />

(Natural History)/Oxford University Press, London,<br />

England. 217 pp.<br />

Rankin, J. S. 1939. The life cycle of the frog bladder<br />

fluke, Gorgoderina attenuata Stafford, 1902<br />

(Trematoda: Gorgoderidae). American Midland<br />

Naturalist 21:476-488.<br />

Scale, D. B. 1987. Amphibia. Pages 4<strong>67</strong>-552 in T. J.<br />

Pandian and F. J. Vanberg, eds. Animal Energetics.<br />

Volume 2. Academic Press, San Diego, California.<br />

Smyth, J. D., and M. M. Smyth. 1980. Frogs As<br />

Host-Parasite Systems. 1. Macmillan, London,<br />

England. 212 pp.<br />

Sokal, R. R., and J. F. Rohlf. 1981. Biometry, 2nd<br />

ed. W. H. Freeman and Company, New York. 859<br />

pp.<br />

Toft, C. A. 1981. Feeding ecology of Panamanian litter<br />

Anurans: patterns in diet and foraging mode.<br />

Journal of Herpetology 15:139-144.<br />

Ulmer, M. J. 1970. Studies on the helminth fauna of


Iowa I. Trematodes of amphibians. American<br />

Midland Naturalist 83:38-64.<br />

-, and H. A. James. 1976. Studies on the helminth<br />

fauna of Iowa II. Cestodes of amphibians.<br />

Proceedings of the Helminthological Society of<br />

Washington 43:191-200.<br />

Vanderburgh, D. J., and R. C. Anderson. 1987. Pre<br />

liminary observations on seasonal changes in<br />

prevalence and intensity of Cosmocercoides variabilis<br />

(Nematoda: Cosmocercoidea) in Bufo<br />

americanus (Amphibia). Canadian Journal of Zoology<br />

65:1666-16<strong>67</strong>.<br />

BOLEK AND COGGINS—HELMINTH COMMUNITIES IN TOADS 209<br />

Meeting Notices<br />

Vogt, R. C. 1981. Natural History of Amphibians and<br />

Reptiles of Wisconsin. The Milwaukee Public<br />

Museum and Friends of the Museum, Inc., Milwaukee,<br />

Wisconsin. 205 pp.<br />

Williams, D. D., and S. J. Taft. 1980. Helminths of<br />

anurans from NW Wisconsin. Proceedings of the<br />

Helminthological Society of Washington 47:278.<br />

Yoder, H. R. 1998. Community ecology of helminth<br />

parasites infecting molluscan and amphibian<br />

hosts. Ph.D. Thesis, University of Wisconsin-Milwaukee.<br />

138 pp.<br />

The 8th European Multicolloquium of <strong>Parasitology</strong> (EMOP-8) will be held on 10-14 September,<br />

<strong>2000</strong> in Poznari, Poland. "<strong>Parasitology</strong> at the Turn of the Millenium" is the theme of the EMOP-<br />

8, and the program will consist of: 6 symposia (Malaria in European Travelers; Congenital Toxoplasmosis;<br />

Toxocara and Toxocariasis; Tapeworm Zoonoses; Human and Animal Trematode Infections;<br />

and the Status of Food-Born Parasites at the Dawn of the Millenium), 8 scientific sessions<br />

(Biology and Taxonomy; Molecular Biology; Immunology, Including Vaccines; Epidemiology and<br />

Control; Parasitic Infections and Diseases; Antiparasitic Drugs and Drug Resistance; The Parasites<br />

of Fishes and Other Hosts from the Aquatic Environment; and General <strong>Parasitology</strong>, including<br />

SNOPAD), and 5 technical workshops (Microscopy Updated; New Methods in Serological Diagnosis<br />

of Parasitic Infections; Molecular Methods in Diagnosis of Parasitic Infections; Education<br />

in <strong>Parasitology</strong> on CD-ROM; and <strong>Parasitology</strong> on the Internet). Contact information: Organizing<br />

Committee (Prof. Z. S. Pawlowski and Prof. K. Boczofi), Department of Biology and Medical<br />

<strong>Parasitology</strong>, Karl Marcinkowski University of Medical Sciences, Fredry Street 10, 61-701 Poznari,<br />

Poland. Telephone/Fax (48-61) 852-71-92, e-mail: emop8@eucalyptus.usoms.poznan.pl. Internet:<br />

http//www.emop8.am.poznan.pl.<br />

The 4th International Symposium on Monogenea will be held on 9 -13 July 2001 at the Women's<br />

<strong>College</strong> of the University of Queensland, Brisbane, Queensland, Australia. A local committee has<br />

been established to organize details for the various scientific sessions, social events, excursions and<br />

accompanying persons program. Subject to sufficient demand, there will be a post-symposium<br />

workshop on Heron Island on the Great Barrier Reef. Calls for expressions of interest to attend the<br />

symposium, and details regarding registration, submission of abstracts, and a preliminary scientific<br />

program will be available later in <strong>2000</strong>. Contact information: Dr. Ian D. Whittington and Dr.<br />

Leslie A. Chisolm, Department of Microbiology and <strong>Parasitology</strong>, The University of Queensland,<br />

Brisbane, Queensland 4072, Australia. Fax +61 7 3365 4620, e-mail: i.wmttmgton@mailbox.uq.<br />

edu.au or l.chisolm@mailbox.uq.edu.au. For further information and notices see the Internet web<br />

page at: http://www.biosci.uq.edu.au/micro/academic/ianw/ism4.htm.<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 210-217<br />

Ecological Aspects of Endohelminths Parasitizing Cichla monoculus<br />

Spix, 1831 (Perciformes: Cichlidae) in the Parana River near Porto<br />

Rico, <strong>State</strong> of Parana, Brazil<br />

PATRICIA MIYUKI MACHADO,1 SILVIA CAROLINA DE ALMEIDA,1<br />

GlLBERTO CEZAR PAVANELLI,2'3 AND RlCARDO MASSATO TAKEMOTO2<br />

1 Postgraduate Course in Ecology of Continental Aquatic Environments and<br />

2 Center for Research in Limnology, Ichthyology and Aquaculture (DBI/NUPELIA), <strong>State</strong> University of<br />

Maringa, Avenida Colombo 5790, Bloco G-90, 87020-900 Maringa, Parana, Brazil (e-mail:<br />

gcpavanelli@uem.br)<br />

ABSTRACT: We examined 136 specimens of Cichla monoculus Spix, 1831, collected in the Parana River near<br />

Porto Rico, <strong>State</strong> of Parana, Brazil, from July 1996 through October 1997. Of the total number of fish, 133<br />

(97.8%) were infected with at least 1 species of helminth. A total of 8 helminth species was recorded: 3 Digenea,<br />

Clinostomum sp., Diplostomum (Austrodiplostomuni) compactum (Lutz, 1928), and Diplostomum sp.; 3 Cestoda,<br />

Proteocephalus microscopicus Woodland, 1935, Proteocephalus rnacrophallus (Diesing, 1850), and Sciadocephalus<br />

megalodiscus Diesing, 1850; 1 Nematoda, Contracaecum sp.; and 1 Acanthocephala, Quadrigyrus machadoi<br />

Fabio, 1983. Proteocephalus microscopicus and P. macrophallus showed the highest values of prevalence<br />

and intensity of infection, followed by Contracaecum sp. In the endoparasite community of C. monoculus, the<br />

cestodes are both dominant and codominant species. The typical pattern of overdispersion or aggregation was<br />

observed for P. microscopicus, P. macrophallus, S. megalodiscus, Q. machadoi, and Contracaecum sp. Prevalence<br />

and total host length were positively correlated in fish parasitized by P. microscopicus, P. macrophallus,<br />

and S. megalodiscus. Infection intensity and host length were positively correlated only for P. microscopicus.<br />

There were significant differences in the prevalence of P. macrophallus and Q. machadoi in males and females<br />

of C. monoculus. Clinostomum sp., Diplostomum sp., D. (A.) compactum, and Q. machadoi were found for the<br />

first time in C. monoculus.<br />

KEY WORDS: ecology, endohelminths, Digenea, Cestoda, Nematoda, Acanthocephala, freshwater fish, tucunare,<br />

Cichla monoculus, Cichlidae, Teleostei, Parana River, Brazil.<br />

Of the main factors influencing the composition<br />

of endoparasite communities, the feeding<br />

habits of the hosts are most important, since diverse<br />

animals that serve as intermediate hosts<br />

for the hosts' parasites are found in their diets<br />

(Dogiel, 1970). Changes in the diet and feeding<br />

habits of fishes also influence the composition<br />

of their parasite fauna (Dogiel, 1970) and account<br />

for the differences in the parasite faunas<br />

of young and adults (Burn, 1980; Scott, 1982;<br />

Moser and Hsieh, 1992). It is also well understood<br />

that the fluctuations in water level characteristic<br />

of floodplains may modify the feeding<br />

habits of fish because of changes in the quantity<br />

and quality of available food (Junk, 1980; Lowe-<br />

McConnell, 1987; Brasil-Sato and Pavanelli,<br />

1999). The influence of the sex of the hosts is<br />

another important factor responsible for the variation<br />

in the composition of their parasitofauna<br />

and may be related to behavioral, biological, and<br />

physiological differences between male and fe-<br />

3 Corresponding author.<br />

210<br />

Copyright © 2011, The Helminthological Society of Washington<br />

male fish (Paling, 1965; Muzzall, 1980; Fernandez,<br />

1985; Moser and Hsieh, 1992; Takemoto et<br />

al., 1996; Machado et al., 1994). The number of<br />

studies in Brazil on the ecology of helminth parasites<br />

of fish, especially in floodplain environments,<br />

is still small.<br />

The tucunare, Cichla monoculus Spix, 1831,<br />

the object of this investigation, is an important<br />

commercial and sport fish in the Upper Parana<br />

River. It is a native of the Amazon Basin and<br />

was first recorded in the Parana River in 1986<br />

(Agostinho et al., 1994). It is a predator (Lowe-<br />

McConnell, 1969), considered piscivorous because<br />

of the predominance of fish in its diet<br />

(FUEM/CIAMB/PADCT, 1995) and carnivorous<br />

because it eats shrimp (Fontenele and Peixoto,<br />

1979; Gery, 1984; Bonetto and Castello, 1985)<br />

and benthopelagic fish (Ortega and Vari, 1986).<br />

This study was intended to extend our existing<br />

knowledge of the helminth fauna parasitizing<br />

the fishes of the Parana River floodplain, to<br />

show the structure and diversity of the endoparasitic<br />

infrapopulations of the tucunare, and to


MACHADO ET AL.—ECOLOGY OF ENDOHELMINTHS OF CICHLA MONOCULUS 211<br />

Table 1. Prevalence (P%), mean intensity of infection (Mil), mean abundance (MA), range of variation<br />

(Rx), and sites of infection of the endohelminths of 136 specimens of Cichla monoculus collected in Pau<br />

Veio Bayou near Porto Rico, state of Parana, Brazil, from July 1996 through October 1997.*<br />

Parasite Ni Np P (%) Mil ± SD MA ± SD Rx Site of infection<br />

Digenea<br />

Clinostomum sp. (M)<br />

D. (A.) compactum (M)<br />

Diplostomum sp. (M)<br />

Cestoda<br />

Proteocephalus<br />

microscopicus (A)<br />

Proteocephalus<br />

macrophallus (A)<br />

Sciadocephahts<br />

megalodiscus (A)<br />

Nematoda<br />

Contracaecum sp. (L)<br />

Acanthocephala<br />

Quadrigyrus machadoi (L)<br />

2 18 1.5 9.0 ± 9.9 0.13 ± 1.4 2-16 Branchial cavity and<br />

stomach<br />

7 19 5.2 2.7 ± 1.7 0.1 ± 0.7 1-5 Vitreous humor (eye)<br />

12 16 8.8 1.3 ± 0.5 0.1 ± 0.4 1-2 Vitreous humor (eye)<br />

128 36,863 94.1 288.0 ± 793.0 27 1.1 ±772.1 1-8,594 Stomach and intestine<br />

61 1,121 44.9 18.4 ± 74.6 8.2 ± 50.6 1-573 Stomach and intestine<br />

18 154 13.2 8.6 ±11.0 1.1+4.9 1-42 Stomach and intestine<br />

96 1,034 70.6 10.8 ± 32.1 7.6 ± 27.3 1-309 Mesentery<br />

30 76 22.1 2.5 ± 2.0 0.6 ±1.4 1-7 Mesentery (encysted<br />

L); stomach and intestine<br />

(free L)<br />

Ni = number of infected fish; Np = number of parasites; M = metacercaria; A = adults; L = larvae.<br />

analyze the possible influences of sex and length<br />

of the hosts on these infrapopulations.<br />

Materials and Methods<br />

The fish were collected monthly in Pau Veio Bayou<br />

on Mutum Island in the floodplain of the Upper Parana<br />

River, state of Parana, Brazil (22°45'00"S,<br />

53°16'50"W) from July 1996 through October 1997.<br />

After capture and identification, the fish were measured,<br />

weighed, and sexed. They were eviscerated, and<br />

the visceral cavity, eyes, digestive tract and associated<br />

organs, kidney, urinary and reproductive tracts, and<br />

gonads were removed. The organs were separated and<br />

placed in Petri dishes containing 0.65% physiological<br />

solution and examined individually with a stereomicroscope.<br />

The digenetic trematodes were compressed<br />

between slides and/or coverglasses and fixed in cold<br />

AFA. The cestodcs and nematodes were fixed in warm<br />

formol. The acanthocephalans were killed in distilled<br />

water in Petri dishes under refrigeration and fixed unpressed<br />

in AFA. All of the worms were preserved in<br />

70% alcohol. The digenetic trematodes, cestodes, and<br />

acanthocephalans were stained with acetic carmine or<br />

Delafield's hematoxylin. All of the worms were dehydrated<br />

in a graded ethanol series, cleared with<br />

beechwood creosote, and mounted in Canada balsam.<br />

For identification of the parasites, the following works<br />

were used: Diesing (1850), LaRue (1914), Woodland<br />

(1933, 1935), Yamaguti (1963), Freze (1965), Travassos<br />

et al. (1969), Schmidt and Hugghins (1973), Moravec<br />

(1994), Rego (1994), Takemoto and Pavanelli<br />

(1996), Sholtz et al. (1996), and Silva-Souza (1998).<br />

Helminths were deposited in the Helminthological<br />

Collection of the Institute Oswaldo Cruz (FIOCRUZ),<br />

Rio de Janeiro, state of Rio de Janeiro, Brazil, under<br />

the following accession numbers: Clinostomum sp.<br />

34235, Diplostomum (Austrodiplostomwri) compactum<br />

34233, Diplostomum sp. 34232, Proteocephalus microscopicus<br />

34234, Proteocephalus macrophallus<br />

34230, S. megalodiscus 33951, 33952, and 33953 ac,<br />

Contracaecum sp. 34231, and Quadrigyrus machadoi<br />

34236.<br />

Parasite diversity was evaluated by the Shannon diversity<br />

index (H'). The possible variation in parasite<br />

diversity was analyzed in relation to sex of the hosts<br />

by Student's Mest, and in relation to the total length<br />

of the hosts by the Spearman rank correlation coefficient<br />

(rs) (Ludwig and Reynolds, 1988). The importance<br />

value (I) proposed by Bush, according to Thul<br />

et al. (1985), was used to classify the parasite community<br />

components. Species in the larval stage were<br />

not considered in this classification. The dispersion index<br />

was used to determine the distribution of the infrapopulation<br />

in the sample. The degree of overdispersion<br />

or aggregation was calculated using Green's<br />

index (Ludwig and Reynolds, 1988). These tests were<br />

applied only to the endohelminth species present at<br />

prevalences higher than 10%. The correlation between<br />

total host length and the intensity of infection of the<br />

parasite species was evaluated by Spearman rank correlation<br />

coefficient (rs) (Zar, 1996). The existence of a<br />

correlation between total host length and prevalence of<br />

infection was tested using Pearson's correlation coefficient<br />

(r) (9 length classes between 13.1 and 49 cm<br />

were established) after angular transformation of the<br />

prevalence data (arc sinVx)(Zar, 1996). Student's r-test<br />

was used to compare the total lengths of male and<br />

Copyright © 2011, The Helminthological Society of Washington


Table 2. Monthly values of prevalence (P%) and mean intensity of infection (Mil) of the endohelminths<br />

in Pau Veio Bayou near Porto Rico, state of Parana, Brazil, from July 1996 through October 1997 (N =<br />

Proteocephalus S<br />

macrophallus m<br />

P(%) Mil P<br />

Proteocephalus<br />

rnicroscopicus<br />

Diplostomurn<br />

sp.<br />

Diplostornum<br />

Clinostomum (A.)<br />

sp. compactum<br />

Mil<br />

P(%)<br />

Mil<br />

P(%)<br />

N P(%) Mil P(%) Mil<br />

Month<br />

4.8 6<br />

2.0<br />

1.0<br />

12.8 5<br />

179.5<br />

53.0 3<br />

50.0<br />

50.0<br />

100.0<br />

71.4<br />

80.0<br />

33.3<br />

195.6<br />

37.0<br />

42.0<br />

448.9<br />

2,562.5<br />

240.0<br />

87.5<br />

50.0<br />

100.0<br />

100.0<br />

80.0<br />

100.0<br />

1.5<br />

2.0<br />

25.0<br />

50.0<br />

8 12.5 16.0 — —<br />

2 — — — —<br />

2<br />

1.3<br />

1.0<br />

—<br />

42.9<br />

20.0<br />

—<br />

5 — — — —<br />

3 — — — —<br />

1996<br />

Jul<br />

Aug<br />

Sep<br />

Oct<br />

Nov<br />

Dec<br />

4.0<br />

1.0<br />

9.8<br />

6.0<br />

1.8<br />

3.3<br />

1.0<br />

2.5<br />

1.0<br />

12.0<br />

100.0<br />

12.5<br />

66.7<br />

46.2<br />

40.0<br />

37.5<br />

42.9<br />

22.2<br />

100.0<br />

100.0<br />

1,074.0<br />

1.0<br />

293.8<br />

335.7<br />

180.2<br />

98.4<br />

131.3<br />

112.4<br />

1 34.0<br />

531.7<br />

100.0<br />

84.4<br />

100.0<br />

100.0<br />

100.0<br />

100.0<br />

100.0<br />

100.0<br />

100.0<br />

100.0<br />

—<br />

—<br />

1.5<br />

1.0<br />

—<br />

1.0<br />

—<br />

—<br />

—<br />

—<br />

—<br />

—<br />

8.3<br />

15.4<br />

—<br />

12.5<br />

—<br />

—<br />

—<br />

—<br />

2 — — 100.0 3.5<br />

32 — — 3.1 1.0<br />

24 4.2 2.0 8.3 2.5<br />

13 — — 7.7 5.0<br />

10 — — 10.0 1.0<br />

8 — — — —<br />

7 — — — —<br />

9 _ _ _ __<br />

1 _ _ _ __<br />

3 — — — —<br />

1997<br />

Jan<br />

Feb<br />

Mar<br />

Apr<br />

May<br />

Jun<br />

Jul<br />

Aug<br />

Sep<br />

Oct<br />

Copyright © 2011, The Helminthological Society of Washington<br />

COMPARATIVE PARASITOLOGY, <strong>67</strong>(2;


40-<br />

35-<br />

30-<br />

25-<br />

20-<br />

15-<br />

10-<br />

5-<br />

n .<br />

18.1<br />

27.1<br />

MACHADO ET AL.—ECOLOGY OF ENDOHELMINTHS OF CICHLA MONOCULUS 213<br />

39.1<br />

1 2 3 4 5<br />

No. parasite species<br />

Figure I. Parasite richness in 136 specimens of<br />

Clchla monoculus collected in Pau Veio Bayou near<br />

Porto Rico, state of Parana, Brazil from July 1996<br />

through October 1997.<br />

female hosts. The effect of sex of the host on the prevalence<br />

of each parasite species was evaluated by the<br />

log-likelihood G test using a 2 X 2 contingency table<br />

(Zar, 1996), and the intensities of infection of each<br />

species of parasite in the male and female hosts were<br />

compared using the Mann-Whitney (/-test (Siegel,<br />

1975). In the data analysis, only values with significance<br />

levels of P •& 0.05 were considered significant.<br />

Ecological terms are based on Bush et al. (1997).<br />

Results<br />

Of the 136 hosts examined, 133 (97.8%) were<br />

parasitized by 1 or more species of endohelminth,<br />

of which 39,301 specimens were collected.<br />

The endohelminths included 3 species of digeneans<br />

(Clinostomum sp., Diplostomum (Austrodiplostomum)<br />

compactum (Lutz, 1928), and<br />

Diplostomum sp.); 3 species of cestodes (Proteocephalus<br />

microscopicus Woodland, 1935,<br />

Proteocephalus macrophallus (Diesing, 1850),<br />

and Sciadocephalus megalodiscus Diesing,<br />

1850); 1 species of nernatode (Contracaecum<br />

sp.); and 1 species of acanthocephalan (Quadrigyrus<br />

machadoi Fabio, 1983) that were found<br />

free in the intestine or encysted in the mesentery<br />

in the same stage of development (Tables 1 and<br />

2). Parasite richness varied from 1 to 5 species,<br />

and 52 fish were infected by 3 species of parasites<br />

(Fig. 1).<br />

The cestodes were the most frequently encountered<br />

parasites, corresponding to 97% of the<br />

helminths collected, and were present in 130<br />

hosts. Proteocephalus microscopicus showed<br />

the highest percentages of parasitism, followed<br />

by Contracaecum sp. (Table 1). Proteocephalus<br />

microscopicus also was the species that presented<br />

the highest monthly values of prevalence and<br />

mean intensity (Table 2). The acanthocephalan<br />

Q. machadoi made up 0.2% of the parasite spec-<br />

11.3<br />

1.5<br />

Table 3. Classification and Bush's Importance values<br />

(I) of the species of endohelminth parasites of<br />

136 specimens of Cichla monoculus collected in Pau<br />

Veio Bayou near Porto Rico, state of Parana, Brazil<br />

from July 1996 through October 1997.<br />

Parasite I<br />

Dominant species<br />

Proteocephalus microscopicus 98.50<br />

Proteocephalus macrophallus 1.40<br />

Co-dominant species<br />

Sciadocephalus megalodiscus 0.06<br />

imens collected, while the digenetic trematodes<br />

represented 0.14%.<br />

The mean parasite diversity according to the<br />

Shannon index (H') was 0.1329 (SD = 0.1879),<br />

and the maximum diversity was 1.3716. Parasite<br />

diversity did not differ significantly between<br />

male and female hosts (t = 0.6004, P = 0.5492),<br />

and was not correlated with total length of the<br />

hosts (rs = 0.03124, P = 0.7180). Total host<br />

length varied from 13.5 to 45.7 cm (mean 25.1<br />

cm).<br />

Using the importance value (I) proposed by<br />

Bush, 2 species were classified as dominants and<br />

1 as co-dominant (Table 3). The parasites of C.<br />

monoculus showed the typical pattern of overdispersion<br />

or aggregation of the parasite populations.<br />

Proteocephalus macrophallus and S.<br />

megalodiscus showed the highest values of<br />

Green's index of aggregation (Table 4). There<br />

was no significant difference in length between<br />

the 61 male and 75 female tucunares examined<br />

(t = 0.6130, P = 0.5409).<br />

There was a positive correlation between total<br />

length of the hosts and prevalence for fish parasitized<br />

by P. microscopicus, P. macrophallus,<br />

and S. megalodiscus. A positive correlation be-<br />

Table 4. Dispersion Index (DI) and Green's Index<br />

of Aggregation (GI) for the endohelminth parasite<br />

species of 136 specimens of Cichla monoculus collected<br />

in Pau Veio Bayou near Porto Rico, state of<br />

Parana, Brazil from July 1996 through October<br />

1997.<br />

Parasite<br />

Proteocephalus microscopicus<br />

Proteocephalus macrophallus<br />

Sciadocephalus megalodiscus<br />

Quadrigyrus machadoi<br />

Contracaecum sp.<br />

Copyright © 2011, The Helminthological Society of Washington<br />

DI<br />

2,198.96<br />

312.23<br />

21.83<br />

3.27<br />

98.06<br />

GI<br />

0.059<br />

0.278<br />

0.136<br />

0.030<br />

0.094


. JI<br />

Table 5. Values of Spearman's rank correlation coefficients (rs) and Pearson's correlation coefficients<br />

(r), to evaluate the relationship between intensity and prevalence of infection, respectively, of the endohelminth<br />

fauna with the total length of 136 specimens of Cichla monoculus collected in Pau Veio Bayou<br />

near Porto Rico, state of Parana, Brazil from July 1996 through October 1997.<br />

Parasite<br />

Proteocephalus microscopicus<br />

Proteocephalus macrophallus<br />

Sciadocephalus megalodiscus<br />

Contracaecum sp.<br />

Quadrigyrus machadoi<br />

Level of significance.<br />

rs<br />

0.3596<br />

0.2175<br />

0.1943<br />

-0.0755<br />

-0.2954<br />

tween the total length of the hosts and the intensity<br />

of parasite infection was seen only for P.<br />

microscopicus (Table 5).<br />

In the cestode P. macrophallus and the acanthocephalan<br />

Q. machadoi, prevalence and intensity<br />

of infection were influenced by sex of the<br />

host (Table 6). For P. macrophallus these indices<br />

were higher in the males, and for Q. machadoi<br />

in the females (Table 6). In addition to these<br />

species, S. megalodiscus showed a significant<br />

difference in intensity of infection according to<br />

sex of the host, with higher intensity in males<br />

(Tables 6 and 7).<br />

Discussion<br />

In the tucunare, the proteocephalid cestodes<br />

P. microscopicus and P. macrophallus showed<br />

the highest values of mean intensity of infection.<br />

This is explainable by the feeding habits of the<br />

tucunare, which includes in its diet fish species<br />

that act as intermediate or paratenic hosts of<br />

these parasites. The cestodes found in the present<br />

work appear to be exclusive to the tucunare,<br />

never having been recorded in other fish (Rego,<br />

1994).<br />

For the acanthocephalans, the main factor regulating<br />

the prevalence and intensity of the par-<br />

P*<br />

P < 0.0001<br />

P = 0.0922<br />

P = 0.4397<br />

p = 0.4648<br />

P = 0.1130<br />

r<br />

0.7501<br />

0.9048<br />

0.7603<br />

0.2509<br />

0.5697<br />

P<br />

P = 0.0199<br />

P = 0.0008<br />

P = 0.0174<br />

P = 0.5149<br />

P = 0.1093<br />

asitoses is also predation by the fish on the intermediate<br />

or paratenic hosts (Amin and Burrows,<br />

1977). In the specimens of C. monoculus<br />

studied, only larvae of Q. machadoi in the same<br />

stage of development were found free in the intestine<br />

or encysted in the mesentery, which may<br />

indicate that these fish are intermediate or paratenic<br />

hosts of this parasite. The natural predators<br />

of tucunare in the study region are carnivorous<br />

fish or piscivorous birds. The small number<br />

of fish infected by Clinostomum sp., together<br />

with the fact that the worms were found free in<br />

the stomach, may indicate that they were ingested<br />

accidentally with prey.<br />

Only 2 individuals of C. monoculus showed<br />

the maximum parasite richness found, i.e., 5<br />

species of endoparasites. Most of the population<br />

was parasitized by 3 species of helminths, of<br />

which P. microscopicus was always present.<br />

Holmes (1990) pointed out that parasite richness<br />

is higher in fishes of intermediate trophic levels,<br />

since they harbor both adult and larval stages of<br />

parasites.<br />

The relationship between body length or age<br />

of the host and parasite diversity is based on the<br />

process of temporal accumulation and on the increase<br />

in the dimensions of the sites of infection<br />

Table 6. Results of the log-likelihood test (G) to compare prevalence between males and females and of<br />

the Mann-Whitney l/-test to compare intensity of infection between males and females of 136 specimens<br />

of Cichla monoculus collected in Pau Veio Bayou near Porto Rico, state of Parana, Brazil from July 1996<br />

through October 1997.<br />

Parasite<br />

Proteocephalus microscopicus<br />

Proteocephalus macrophallus<br />

Sciadocephalus megalodiscus<br />

Contracaecum sp.<br />

Quadrigyrus machadoi<br />

G<br />

1.0<strong>67</strong><br />

5.321<br />

2.210<br />

0.111<br />

5.410<br />

P<br />

0.5 > P > 0.25<br />

0.025 > P > 0.01<br />

0.25 > P > 0.10<br />

0.75 > P > 0.50<br />

0.025 > P > 0.01<br />

P = level of significance; Z = value of normal approximation of {/-test.<br />

Z<br />

0.16<br />

7.32<br />

3.18<br />

1.05<br />

5.71<br />

Copyright © 2011, The Helminthological Society of Washington<br />

P<br />

P > 0.25<br />

P < 0.0005<br />

0.01 > P > 0.005<br />

0.25 > P > 0.10<br />

P < 0.005


MACHADO ET AL.—ECOLOGY OF ENDOHELMINTHS OF CICHLA MONOCULUS 215<br />

Table 7. Prevalence and mean intensity of infection by helminth parasites of Cichla monoculus collected<br />

in Pau Veio Bayou near Porto Rico, state of Parana, Brazil from July 1996 through October 1997.*<br />

Parasite Nm Nf Nmi Nfi Pm (%) Pf (%) Mim Mif<br />

Digenea<br />

Clinostomum sp.<br />

Diplostomum (A.) compactum<br />

Diplostomiim sp.<br />

Cestoda<br />

Proteocephalus microscopicus<br />

Proteocephalus macrophallus<br />

Sciadocephalus megalodiscus<br />

Nematoda<br />

Contracaecum sp.<br />

Acanthocephala<br />

Quadrigyrus machadoi<br />

61<br />

61<br />

61<br />

61<br />

61<br />

61<br />

61<br />

61<br />

75<br />

75<br />

75<br />

75<br />

75<br />

75<br />

75<br />

75<br />

0<br />

2<br />

6<br />

56<br />

34<br />

11<br />

44<br />

8<br />

2<br />

5<br />

6<br />

72<br />

27<br />

7<br />

52<br />

22<br />

0.0<br />

3.3<br />

9.8<br />

91.8<br />

55.7<br />

18.0<br />

70.5<br />

13.1<br />

2.7<br />

6.7<br />

8.0<br />

96.0<br />

36.0<br />

9.3<br />

70.7<br />

29.3<br />

—<br />

4.0 ± 0.0<br />

1.2 ± 0.4<br />

434.7 ± 1,160.2<br />

26.4 ± 99.0<br />

9.3 ± 12.0<br />

14.5 ± 46.4<br />

2.0 ±<br />

1.1<br />

9.0 ± 9.9<br />

2.2 ± 1.8<br />

1.5 ± 0.6<br />

173.9 ± 227.6<br />

8.4 ± 15.3<br />

7.4 ± 10.0<br />

7.7 ± 8.8<br />

2.7 ± 2.2<br />

* Nm = number of males examined; Nf = number of females examined; Nmi = number of males infected; Nfi = number of<br />

females infected; Pm and Pf = prevalence of males and females, respectively; Mim and Mif = mean intensity of infection of<br />

males and females, respectively.<br />

as a function of growth (Luque et al., 1996).<br />

Such a relationship has been shown not to exist<br />

in other species of freshwater fishes (Adams,<br />

1986; Janovy and Hardin, 1988; Machado et al.,<br />

1996). The lack of this relationship in the tucunare<br />

may indicate a homogeneity in their<br />

feeding habits during ontogenetic development.<br />

The independence of diversity in relation to<br />

the sex of C. monoculus may constitute evidence<br />

that the occupation of the habitat and the diet<br />

are similar in males and females. Adams (1986),<br />

Janovy and Hardin (1988), and Machado et al.<br />

(1996) obtained similar results for other species<br />

of freshwater fishes.<br />

The cestodes P. microscopicus, P. macrophallus,<br />

and S. megalodiscus must be considered<br />

as basic components of the parasite community<br />

of C. monoculus. The first 2 species were classified<br />

as dominants, and the third as codominant<br />

(Thul et al., 1985). The dominant and codominant<br />

species of endohelminths showed a pattern<br />

of spatial aggregation, in agreement with the<br />

typical pattern of endoparasitism demonstrated<br />

by other investigators (Skorping, 1981; Janovy<br />

and Hardin, 1987; Oliva et al., 1990; Takemoto,<br />

1993; Machado et al., 1996).<br />

The positive correlation observed between the<br />

total length of the hosts and the prevalence of<br />

P. microscopicus, P. macrophallus, and S. megalodiscus<br />

indicates the occurrence of a cumulative<br />

process. A positive correlation between<br />

the standard length of hosts and the prevalence<br />

and/or intensity of infection was observed also<br />

by Conneely and McCarthy (1986), Machado et<br />

al. (1994), and Takemoto and Pavanelli (1994);<br />

the latter 2 investigations were also carried out<br />

in the Parana River.<br />

Esch et al. (1988) pointed out that the sex of<br />

the hosts may also be one of the factors that<br />

influence levels of parasitism. The influence of<br />

physiological factors (hormones, mucosity) was<br />

demonstrated by Paling (1965) and Moser and<br />

Hsieh (1992), who hypothesized that certain<br />

species of parasites possess a greater facility to<br />

infect male or female hosts. Muzzall (1980) and<br />

Takemoto and Pavanelli (1994) found no influence<br />

of the sex of the host on the parasite fauna,<br />

showing that the ecological relationships (behavior,<br />

habitat, and diet) of males and females<br />

are similar. In C. monoculus, the sex influences<br />

the prevalence and intensity of infection of P.<br />

macrophallus and Q. machadoi and only the intensity<br />

of S. megalodiscus. This can be explained<br />

by the fact that some species of fish become<br />

more susceptible in the breeding season<br />

because of physiological and behavioral chang-<br />

Acknowledgments<br />

We are grateful to Drs. Janet W. Reid and Willis<br />

A. Reid, Jr., who assisted in translating the<br />

text into English.<br />

Copyright © 2011, The Helminthological Society of Washington


216 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Literature Cited<br />

Adams, A. M. 1986. The parasite community on the<br />

gills of Fundulus kansae (Garman) from the South<br />

Platte River, Nebraska (USA). Acta Parasitologica<br />

Polonica 31:47-54.<br />

Agostinho, A. A., H. F. Julio, Jr., and M. Petrere,<br />

Jr. 1994. Itaipu Reservoir (Brazil): impacts of the<br />

impoundment on the fish fauna and fisheries. Pages<br />

171-184 in I. G. Cowx, ed. Rehabilitation of<br />

Freshwater Fisheries. Hull International Fisheries<br />

Institute, University of Hull, U.K.<br />

Amin, O. M., and J. M. Burrows. 1977. Host and<br />

seasonal associations of Echinorhynchus salmonis<br />

(Acanthocephala, Echinorhynchidae) in Lake<br />

Michigan fishes. Journal of the Fisheries Research<br />

Board of Canada 34:325-331.<br />

Bonetto, A. A., and H. P. Castello. 1985. Pesca y<br />

Piscicultura en Aguas Continentales de America<br />

Latina. Secretaria General de la Organizacion de<br />

los Estados Americanos, Programa Regional de<br />

Desarrollo Cientifico y Tecnologico, Washington,<br />

D.C. Monografia 31. 118 pp.<br />

Brasil-Sato, M. C., and G. C. Pavanelli. 1999. Ecological<br />

and reproductive aspects of Neoechinorhynchus<br />

pimelodi Brasil-Sato & Pavanelli (Eoacanthocephala,<br />

Neoechinorhynchidae) of Pirnelodus<br />

maculatus Lacepede (Siluroidei, Pimelodidae)<br />

of the Sao Francisco River, Brasil. Revista Brasileira<br />

de Zoologia 16:73-82.<br />

Burn, P. R. 1980. The parasites of smooth flounder,<br />

Liopsetta putnami (Gill), from the Great Bay Estuary,<br />

New Hampshire. Journal of <strong>Parasitology</strong><br />

66:532-541.<br />

Bush, A. O., K. D. Lafferty, J. M. Lotz, and A. W.<br />

Shostak. 1997. <strong>Parasitology</strong> meets ecology on its<br />

own terms: Margolis et al. revisited. Journal of<br />

<strong>Parasitology</strong> 83:575-583.<br />

Conneely, J. J., and T. K. McCarthy. 1986. Ecological<br />

factors influencing the composition of the parasite<br />

fauna of the European eel, Anguilla anguilla<br />

(L.), in Ireland. Journal of Fish Biology 28:207-<br />

219.<br />

Diesing, K. M. 1850. Systema Helminthum. Vol. 1.<br />

Wilhelm Braumuller, Vindobonae. <strong>67</strong>9 pp.<br />

Dogiel, V. A. 1970. Ecology of the parasites of freshwater<br />

fishes. Pages 1-47 in V. A. Dogiel, G. K.<br />

Petrushevsky, and Yu. I. Polyanski, eds. <strong>Parasitology</strong><br />

of Fishes. T. F. H. Publications, Inc. Ltd.,<br />

Hong Kong.<br />

Esch, G. W., C. R. Kennedy, A. O. Bush, and J. M.<br />

Aho. 1988. Patterns in helminth communities in<br />

freshwater fish in Great Britain: alternative strategies<br />

for colonization. <strong>Parasitology</strong> 96:519-532.<br />

Fernandez, J. 1985. Estudio parasitologico de Merlucclus<br />

australis (Hutton, 1872) (Pisces: Merluccidae):<br />

aspectos sistematicos, estadisticos y zoogeograficos.<br />

Boletin de la Sociedad de Biologfa<br />

de Concepcion 56:31-41.<br />

Fontenele, O., and J. T. Peixoto. 1979. Aprecia9ao<br />

sobre os resultados da introdu9ao do tucunare<br />

comum, Cichla ocellaris (Bloch & Schneider,<br />

1801), nos acudes do nordeste brasileiro, atraves<br />

da pesca comercial. Boletim Tecnico DNOCS,<br />

Fortaleza 37:109-134.<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Freze, V. I. 1965. Principles of Cestodology. K. I.<br />

Skrjabin, ed. Proteocephalata Cestodes of Fishes,<br />

Amphibians and Reptiles. Moscow: Translated<br />

from Russian, Israel Program for Scientific Translations,<br />

Jerusalem. 538 pp.<br />

FUEM/CIAMB/PADCT. 1995. Estudos ambientais<br />

da plamcie de inunda£ao do Rio Parana no trecho<br />

compreendido entre a foz do rio Paranapanema e<br />

o reservatorio de Itaipu. Final Report 2:411—426.<br />

Gery, J. 1984. The fishes of Amazonia. In H. Sioli,<br />

ed. The Amazon: Limnology and Landscape Ecology<br />

of a Mighty Tropical River and Its Basin.<br />

Monographiae Biologicae 56, Dr. W. Junk Publishers,<br />

Dordrecht, The Netherlands.<br />

Holmes, J. C. 1990. Competition, contacts, and other<br />

factors restricting niches of parasitic helminths.<br />

Annales de Parasitologie Humaine et Comparee<br />

65:69-72.<br />

Janovy, J., and E. L. Hardin. 1987. Population dynamics<br />

of the parasites in Fundulus zebrinus in<br />

the Platte River of Nebraska. Journal of <strong>Parasitology</strong><br />

73:689-696.<br />

, and . 1988. Diversity of the parasite<br />

assemblage of Fundulus zebrinus in the Platte<br />

River of Nebraska. Journal of <strong>Parasitology</strong> 74:<br />

207-213.<br />

Junk, W. J. 1980. Areas inundaveis—um desafio para<br />

a limnologia. Acta Amazonica 10:775-796.<br />

LaRue, G. R. 1914. A revision of the cestode family<br />

Proteocephalidae. III. Biological Monographs 1:<br />

1-350.<br />

Lowe-McConnell, R. H. 1969. Speciation in tropical<br />

freshwater fishes. Biological Journal of the Linnean<br />

Society 1:51-75.<br />

. Ecological Studies in Tropical Fish Communities.<br />

Cambridge University Press, Cambridge,<br />

U. K. 382 pp.<br />

Ludwig, J. A., and J. F. Reynolds. 1988. Statistical<br />

Ecology: Primer on Methods and Computing. Wiley-Interscience<br />

Publications, New York. 337 pp.<br />

Luque, J. L., J. F. R. Amato, and R. M. Takemoto.<br />

1996. <strong>Comparative</strong> analysis of the communities of<br />

metazoan parasites of Orthopristis ruber and Haemulon<br />

steindachneri (Osteichthyes: Haemulidae)<br />

from the Southeastern Brazilian littoral: II. Diversity,<br />

interspecific associations, and distribution of<br />

gastrointestinal parasites. Revista Brasileira de<br />

Biologia 56:293-302.<br />

Machado, M. H., G. C. Pavanelli, and R. M. Takemoto.<br />

1994. Influence of host's sex and size on<br />

endoparasitic infrapopulations of Pseudoplatystoma<br />

corruscans and Schizodon borelli (Osteichthyes)<br />

of the High Parana River, Brazil. Revista<br />

Brasileira de Parasitologia Veterinaria 3:143-148.<br />

, , and . 1996. Structure and diversity<br />

of endoparasitic infracommunities and the<br />

trophic level of Pseudoplatystorna corruscans and<br />

Schizodon borelli (Osteichthyes) of High Parana<br />

River. Memorias do Institute Oswaldo Cruz 97:<br />

441-448.<br />

Moravec, F. 1994. Parasitic Nematodes of Freshwater<br />

Fishes of Europe. Kluwer Academic Publishers.<br />

Ceske Budejovice. 473 pp.<br />

Moser, M., and J. Hsieh. 1992. Biological tags for


MACHADO ET AL.—ECOLOGY OF ENDOHELMINTHS OF CICHLA MONOCULUS 217<br />

stock separation in Pacific herring Clupea harengus<br />

pallasi in California. Journal of <strong>Parasitology</strong><br />

78:54-60.<br />

Muzzall, P. M. 1980. Population biology and hostparasite<br />

relationships of Triganodistomum attenuatum<br />

(Trematoda: Lissorchiidae) infecting the<br />

white sucker, Catostornus commersoni (Lacepede).<br />

Journal of <strong>Parasitology</strong> 66:293-298.<br />

Oliva, M., J. L. Luque, and J. A. lannacone. 1990.<br />

The metazoan parasites of Stellifer minor (Tschudi,<br />

1844): an ecological approach. Memorias do<br />

Instituto Oswaldo Cruz 85:271-274.<br />

Ortega, H., and P. R. Vari. 1986. Annotated checklist<br />

of the freshwater fishes of <strong>Peru</strong>. Smithsonian Contributions<br />

to Zoology 437:1-25.<br />

Paling, J. E. 1965. The population dynamics of the<br />

monogenean gill parasite Discocotyle sagittata<br />

Leuckart on Windermere trout, Salmo trutta L.<br />

<strong>Parasitology</strong> 55:6<strong>67</strong>-694.<br />

Rego, A. A. 1994. Order Proteocephalidea Mola,<br />

1928. Pages 257-293 in L. F. Khalil, A. Jones,<br />

and R. A. Bray, eds. Keys to the Cestode Parasites<br />

of Vertebrates. CAB International, Wallingford,<br />

U.K.<br />

Schmidt, G. D., and E. J. Hugghins. 1973. Acanthocephala<br />

of South American fishes. Part 1. Eoacanthocephala.<br />

Journal of <strong>Parasitology</strong> 59:829-835.<br />

Scott, J. S. 1982. Digenean parasite communities in<br />

flatfishes of the Scotia Shelf and southern Gulf of<br />

St. Lawrence. Canadian Journal of Zoology 60:<br />

2804-2811.<br />

Scholz, T., A. Chambrier, A. Prouza, and R. Royero.<br />

1996. Redescription of Proteocephalus macrophallus,<br />

a parasite of Cichla ocellaris (Pisces:<br />

Cichlidae) from South America. Folia Parasitologica<br />

43:287-291.<br />

Siegel, S. 1975. Estatfstica Nao Parametrica (Para<br />

Ciencias do Comportamento). McGraw-Hill do<br />

Brasil, Sao Paulo, Brazil. 350 pp.<br />

Silva-Souza, A. T. 1998. Estudo do Parasitismo de<br />

Plagioscion squamosissimus (Heckel, 1840) (Perciformes,<br />

Scianidae) por Diplostomum (Austrodiplostomum)<br />

compactum (Lutz, 1928) (Trematoda,<br />

Digenea) no Rio Tibagi, PR. Ph.D. Thesis, Universidade<br />

Federal de Sao Carlos. 125 pp.<br />

Skorping, A. 1981. Seasonal dynamics in abundance,<br />

development and pattern of infection of Bunodera<br />

luciopercae (Miiller) in perch, Perca fluviatilis L.<br />

from an oligotrophic lake in Norway. Journal of<br />

Fish Biology 18:401-410.<br />

Takemoto, R. M. 1993. Estudo taxonomico, sistematico<br />

e ecologico dos metazoarios parasites de<br />

Oligoplites palometa, O. saums e O. saliens (Osteichthyes:<br />

Carangidae) do literal do Estado do<br />

Rio de Janeiro. M.S. Dissertation, Universidade<br />

Federal Rural do Rio de Janeiro. 157 pp.<br />

, J. F. R. Amato, and J. L. Luque. 1996.<br />

<strong>Comparative</strong> analysis of the metazoan parasite<br />

communities of leather] ackets, Oligoplites palometa,<br />

O. saums, and O. saliens (Osteichthyes:<br />

Carangidae) from Sepetiba Bay, Rio de Janeiro,<br />

Brazil. Revista Brasileira de Biologia 56:639—<br />

650.<br />

, and G. C. Pavanelli. 1994. Ecological aspects<br />

of proteocephalidean cestodes parasites of<br />

Paulicea luetkeni (Steindachner) (Osteichthyes:<br />

Pimelodidae) from the Parana River, Parana, Brazil.<br />

Revista Unimar 16:17-26.<br />

, and . 1996. Proteocephalidean cestodes<br />

in the freshwater fish Cichla monoculus<br />

from the Parana River, Brazil. Studies on Neotropical<br />

Fauna and Environment 31:123-127.<br />

Thul, J. E., D. J. Forrester, and C. L. Abercrombie.<br />

1985. Ecology of parasitic helminths of wood<br />

ducks Aix sponsa, in the Atlantic Fly way. Proceedings<br />

of the Helminthological Society of<br />

Washington 52:297-310.<br />

Travassos, L., J. F. T. Freitas, and A. Kohn. 1969.<br />

Trematodeos do Brasil. Memorias do Instituto Oswaldo<br />

Cruz <strong>67</strong>:1-886.<br />

Woodland, W. N. F. 1933. On the anatomy of some<br />

fish cestodes described by Diesing from the Amazon.<br />

Quarterly Journal of Microscopical Science<br />

76:175-208.<br />

. 1935. Some new Proteocephalids and a Pythobothriid<br />

(Cestoda) from the Amazon. Proceedings<br />

of the Zoological Society of London 3:619-<br />

624.<br />

Yamaguti, S. 1963. Systema Helminthum. V. Acanthocephala.<br />

Interscience Publishers, New York.<br />

860 pp.<br />

Zar, J. H. 1996. Biostatistical Analysis, 3rd ed. Prentice-Hall,<br />

Inc., Upper Saddle River, New Jersey.<br />

662 pp.<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 218-223<br />

Prevalence of Hookworms (Uncinaria lucasi Stiles) in Northern Fur<br />

Seal (Callorhinus ursinus Linnaeus) Pups on St. Paul Island, Alaska,<br />

U.S.A.: 1986-1999<br />

EUGENE T. LYONS,u TERRY R. SPRAKER,2 KIMBERLY D. OLSON,2 SHARON C. TOLLIVER,'<br />

AND HEATHER D. BAIR'<br />

1 Department of Veterinary Science, University of Kentucky, Gluck Equine Research Center, Lexington,<br />

Kentucky 40546-0099, U.S.A. (e-mail: elyonsl@pop.uky.edu; sctolll@pop.uky.edu) and<br />

2 Department of Pathology, <strong>College</strong> of Veterinary Medicine, Colorado <strong>State</strong> University, Ft. Collins, Colorado<br />

80523, U.S.A. (e-mail: tspraker@vth.colostate.edu)<br />

ABSTRACT: Prevalence of hookworms (Uncinaria lucasi) was studied in northern fur seal (Callorhinus ursinus)<br />

pups at necropsy on St. Paul Island, Alaska, U.S.A. Gross examination of 2,121 pups during the period 1986-<br />

1998 revealed that only 13 had a light hookworm infection and 2 other pups had hookworm disease (moderate<br />

hookworm infection, enteritis, and anemia). Specific parasitologic inspections of 230 pups in 1988 and 1996-<br />

1999 included examinations of feces for eggs or intestines for adult worms. Hookworm eggs were present in<br />

fecal samples of 1 (2%) of 48 pups in 1988 and none of 19 pups in 1996. Specimens of U. lucasi, found mostly<br />

during qualitative examinations, were in the intestines of 4 (5%) of 77 pups in 1997, 4 (9%) of 47 in 1998, and<br />

1 (3%) of 39 in 1999. The 9 infected pups harbored 1-8 hookworms each. Observations in this study indicate<br />

a dramatic decline in hookworm prevalence in C. ursinus pups on St. Paul Island compared with that of several<br />

years previously.<br />

KEY WORDS: Uncinaria lucasi, adult hookworms, prevalence, northern fur seal pups, Callorhinus ursinus, St.<br />

Paul Island, Alaska, U.S.A.<br />

Uncinaria lucasi Stiles, 1901, and Uncinaria<br />

hamiltoni Baylis, 1933, are the only 2 species of<br />

hookworms described from pinnipeds (Baylis,<br />

1933, 1947; Stiles, 1901). However, there are<br />

types with measurements intermediate between<br />

U. lucasi and U. hamiltoni (Baylis, 1933; Dailey<br />

and Hill, 1970). Specimens of Uncinaria are<br />

common in otariids (eared seals) and rare in<br />

phocids (earless or true seals) (George-Nascimento<br />

et al., 1992). Classification of hookworms<br />

in pinnipeds remains uncertain. George-Nascimento<br />

et al. (1992) considered all hookworms in<br />

otariids to be the same species, U. lucasi. In the<br />

present paper, the hookworms are designated as<br />

U. lucasi because they were first described and<br />

named from northern fur seals (Callorhinus ursinus<br />

Linnaeus, 1758). This research was done<br />

to compare the current prevalence of adult U.<br />

lucasi in northern fur seal pups on St. Paul Island,<br />

Alaska, U.S.A., with prevalences found in<br />

earlier studies.<br />

Materials and Methods<br />

Dead fur seal pups were collected from 2 rookeries,<br />

Northeast Point and Reef, on St. Paul Island, Alaska<br />

(57°09'N, 170°13'W), for pathologic studies, from<br />

3 Corresponding author.<br />

218<br />

Copyright © 2011, The Helminthological Society of Washington<br />

1986 to 1999. Some of these pups were selected for<br />

specific research on hookworms. The pups were gathered<br />

daily from early July through the first 2 weeks of<br />

August of each year. Pups selected had not been dead<br />

for more than about 24 hr. Collection was by a person<br />

working on a catwalk over a rookery and using a 5m-long<br />

pole equipped with a noose. Pup carcasses recovered<br />

from rookeries were taken directly to the research<br />

laboratory on St. Paul Island for necropsy.<br />

Gross examination of 2,121 dead pups, for the overall<br />

period from 1986 to 1998, included observations<br />

for hookworms in the opened ileocecal area. About<br />

95% of the pups were from Northeast Point and 5%<br />

from Reef rookeries. The infections were separated<br />

into 2 categories: 1) light hookworm infections, when<br />

only a few parasites were noted, and 2) hookworm<br />

disease, defined as moderate hookworm infection,<br />

when a number of parasites, enteritis, and anemia were<br />

evident.<br />

Special parasitologic examinations of 230 pups were<br />

conducted. Fecal samples were obtained from the colons<br />

of 48 pups in 1988 and 19 pups in 1996, placed<br />

in glass vials or plastic bags containing 5% formalin,<br />

and examined for hookworm eggs (Lyons et al., 1976).<br />

Determinations (Lyons, 1963; Olsen and Lyons, 1962,<br />

1965) for presence of adult hookworms in the intestines<br />

were made for 77 pups in 1997, for 47 in 1998,<br />

and for 39 in 1999. Parts of the intestines examined<br />

were 1) approximately 60-100 mm of the ileum, the<br />

entire cecum, and about 45-60 mm of the proximal<br />

end of the colon for all 77 pups in 1977; 2) up to about<br />

300 mm of the ileum, the entire cecum, and approximately<br />

150 mm of the colon, including the proximal<br />

end, for 36 pups, and the entire intestinal tract for 11


Table 1. Summary of results of gross examination<br />

of the ileocecal area of the intestines of northern<br />

fur seal pups (n = 2,121) for hookworm (Uncinaria<br />

lucasi) infection and hookworm disease at necropsy<br />

from 1986 through 1998 on St. Paul Island, Alaska,<br />

U.S.A.<br />

Year<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

1994<br />

1995<br />

1996<br />

1997<br />

1998<br />

Total<br />

Examined<br />

39<br />

90<br />

91<br />

113<br />

364<br />

248<br />

227<br />

141<br />

251<br />

130<br />

172<br />

165<br />

90<br />

2,121<br />

No. of pups<br />

With light<br />

hookworm<br />

infection<br />

0<br />

5<br />

1<br />

2<br />

0<br />

1<br />

2<br />

0<br />

0<br />

0<br />

0<br />

1<br />

1<br />

13<br />

LYONS ET AL.—HOOKWORMS IN NORTHERN FUR SEAL PUPS 219<br />

With<br />

hookworm<br />

disease*<br />

0<br />

0<br />

0<br />

1<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

1<br />

0<br />

0<br />

2<br />

Table 2. Results of examination of intestines* of<br />

northern fur seal pups at necropsy on St. Paul Island,<br />

Alaska, U.S.A., for the hookworm, Uncinaria<br />

lucasi, in 1997, 1998, and 1999.<br />

Year<br />

1997<br />

1998<br />

1999<br />

Total<br />

No. of pups<br />

Examined<br />

77<br />

47<br />

39<br />

163<br />

Infected<br />

(%)<br />

4(5)<br />

4 (9)<br />

1 (3)<br />

9 (6)<br />

No. (mean) of hookworms/<br />

infected pups<br />

Male Female Total<br />

0.50<br />

0.75<br />

1.00<br />

0.<strong>67</strong><br />

0.75<br />

2.00<br />

0.00<br />

1.38<br />

1.25<br />

2.75<br />

1.00<br />

1.89<br />

* Ileocecal areas for 113 pups and complete intestinal tracts<br />

for 50 pups.<br />

* Moderate hookworm infection, enteritis, and anemia.<br />

Discussion<br />

In this study, prevalence of U. lucasi in northern<br />

fur seal pups on St. Paul Island was low.<br />

This was determined mostly from gross examination<br />

of the ileocecal area of the pups' intestines.<br />

The observed prevalence might have been<br />

higher if the entire small and large intestine had<br />

been scrutinized for every pup. However, Olsen<br />

(1958) demonstrated that U. lucasi concentrate<br />

in the ileum, cecum, and proximal colon; these<br />

pups in 1998; and 3) the entire intestinal tract for 39<br />

pups in 1999.<br />

areas were examined in all fur seal pups in the<br />

present study.<br />

Current rates of infection of adult U. lucasi in<br />

Results<br />

northern fur seal pups appear to be much lower<br />

Results of gross examination of the entire<br />

sample of 2,121 pups at necropsy from 1986 to<br />

1998 are summarized in Table 1. Only 13 pups<br />

had light hookworm infections, and 2 other pups<br />

were considered to have hookworm disease.<br />

Examinations of fecal samples revealed hookworm<br />

eggs in only 1 (2%) of 48 pups in 1988<br />

and none of 19 pups in 1996. Prevalence of U.<br />

lucasi, based on recovery of adult specimens in<br />

the intestines of pups at necropsy, was


220 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Table 3. Prevalence of Uncinaria lucasi in intestines of northern fur seal pups at necropsy in some<br />

previous surveys on St. Paul Island, Alaska, U.S.A.<br />

Reference Year of study<br />

Lucas (1899)* 1897<br />

Olsen (1954) 1952<br />

Olsen (1954) 1953<br />

Olscn (1956, 1958) 1955<br />

Lyons and Olsen (1960) 1960<br />

Rookery<br />

Gorbatch<br />

Kitovi<br />

Lagoon<br />

Lukanin<br />

Northeast Point<br />

Polovina<br />

Reef<br />

Tolstoi<br />

Zapadni<br />

Total<br />

Unknown<br />

Polovina<br />

Kitovi<br />

Polovina<br />

Reef<br />

Tolstoi<br />

Vostochnif<br />

Zapadni<br />

Total<br />

Little Polovina<br />

Polovina<br />

Reef<br />

Vostochnit<br />

Zapadni<br />

Total<br />

No. of pups<br />

Examined Infected (%)<br />

33<br />

17<br />

4<br />

12<br />

10<br />

10<br />

57<br />

109<br />

93<br />

345<br />

42<br />

26<br />

28<br />

164<br />

4<br />

100<br />

112<br />

100<br />

553<br />

30<br />

112<br />

63<br />

30<br />

17<br />

252<br />

* These data apparently are for causes of death of pups due to U. lucasi rather than actual prevalence.<br />

t A rookery on Northeast Point.<br />

rain. The ground where fur seals now breed on<br />

the 2 rookeries (Northeast Point and Reef) in the<br />

present study is generally rocky. Previously,<br />

when populations of fur seals were much higher,<br />

breeding animals were more widely dispersed to<br />

include sandy surfaces on these rookeries<br />

(E.T.L., personal observation). Possibly the decline<br />

in numbers of animals breeding on sandy<br />

areas has contributed to the dramatic decrease in<br />

prevalence of U. lucasi.<br />

Table 4 summarizes literature on prevalences<br />

of adult U. lucasi in northern fur seal pups in<br />

some localities other than St. Paul Island. On the<br />

Commander Islands (Bering Island and Medny<br />

Island), Russia, and the Channel Islands (San<br />

Miguel Island), California, U.S.A., the hookworm<br />

prevalence is currently high, based on examinations<br />

of relatively small numbers of dead<br />

pups.<br />

No definite cause has been determined for the<br />

spectacular decline of hookworm infections in<br />

northern fur seal pups on St. Paul Island. Per-<br />

Copyright © 2011, The Helminthological Society of Washington<br />

15 (45)<br />

7 (41)<br />

0(0)<br />

7 (58)<br />

7 (70)<br />

6 (60)<br />

12 (21)<br />

52 (48)<br />

38(41)<br />

144 (42)<br />

38 (90)<br />

24 (92)<br />

5 (18)<br />

120 (73)<br />

13 (27)<br />

73 (73)<br />

89 (79)<br />

<strong>67</strong> (<strong>67</strong>)<br />

3<strong>67</strong> (66)<br />

20 (<strong>67</strong>)<br />

86 (77)<br />

35 (56)<br />

22 (73)<br />

1 1 (65)<br />

174 (69)<br />

haps it is related to one or more unknown factors<br />

in combination with a corresponding decline in<br />

the herd. Numbers of fur seals in the 20th century<br />

peaked in the 1950s and 1960s and began<br />

to decline in the 1970s (Trites, 1992). Estimated<br />

size of the fur seal population on the Pribilof<br />

Islands (St. Paul Island and St. George Island)<br />

was about 1.5 million in the 1960s (Baker, 1957;<br />

Riley, 1961). This population is now estimated<br />

at about 973,000 (York et al., <strong>2000</strong>). The overall<br />

population decreased about 40% in the last 40<br />

years, but the number of pups born declined<br />

much more; i.e., about 60% or from about<br />

500,000 to 200,000 (York et al., <strong>2000</strong>).<br />

Although it is known that parasitic third-stage<br />

larvae (L3) of U. lucasi can live for many years<br />

in the tissues of northern fur seals, there is no<br />

definite information on the time period or number<br />

of lactations for clearance of these larvae<br />

through the mammary system. In experimental<br />

infections of the nematode Strongyloides ransomi<br />

Schwartz and Alicata, 1930, in pigs, colos-


LYONS ET AL.—HOOKWORMS IN NORTHERN FUR SEAL PUPS 221<br />

Table 4. Prevalence of Vncinaria lucasi in intestines of northern fur seal pups at necropsy in some studies<br />

in Russia and in California, U.S.A.<br />

Location Reference<br />

Russia<br />

Commander Islands<br />

Bering Island<br />

Northern rookery<br />

Northwestern rookery<br />

Medny Island<br />

Southeastern rookery<br />

Urilie rookery<br />

Bering Island<br />

Unknown rookery<br />

California, U.S.A.<br />

Channel Islands<br />

San Miguel Island<br />

West Cove rookery<br />

Adams Cove rookery<br />

Year of<br />

study<br />

Kolevatova et al. (1978) 1997<br />

Mizuno (1997)<br />

Lyons et al. (1997)<br />

trum samples were collected for 4 consecutive<br />

lactations from 6 individual sows (Stewart et al.,<br />

1976). There was an exponential decline of parasitic<br />

L3 of S. ransomi with each lactation. The<br />

mean number of larvae/ml of colostrum was 1.1<br />

for the first lactation and 0.06 for the fourth lactation.<br />

Several aspects of the life cycle of U. lucasi,<br />

as studied in northern fur seals on St. Paul Island<br />

(Lyons, 1994), may have a role in the decline in<br />

prevalence: 1) the only source of adult hookworms<br />

in pups is from parasitic L3 passed in the<br />

mother's milk for


222 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

ever, at the present time on St. Paul Island,<br />

hookworm infections and numbers of infected<br />

fur seals are so low that these parasites appear<br />

to be at a minimal or subminimal level for continued<br />

existence. If older fur seals had infections<br />

of adult hookworms contributing eggs to the environment,<br />

continued success of the hookworms<br />

presumably would be more likely. Possibly the<br />

highly evolved, exclusive manner of transmammary<br />

transmission of U. lucasi has become a<br />

detriment for the species under the present conditions<br />

on St. Paul Island.<br />

Acknowledgments<br />

This investigation was made in connection<br />

with a project of the Kentucky Agricultural Experiment<br />

Station and is published with the approval<br />

of the director as paper No. 99-14-5. The<br />

research was conducted under Marine Mammal<br />

Protection Act Research Permit No. 837, issued<br />

to the National Marine Mammal Laboratory, Seattle,<br />

Washington, U.S.A.<br />

Literature Cited<br />

Baker, R. C. 1957. Fur Seals of the Pribilof Islands.<br />

Conservation in Action, Number 12. Fish and<br />

Wildlife Service, U.S. Department of the Interior,<br />

Washington, D.C. 23 pp.<br />

Baylis, H. A. 1933. A new species of the nematode<br />

genus Uncinaria from a sea-lion, with some observations<br />

on related species. <strong>Parasitology</strong> 25:<br />

308-316.<br />

. 1947. A redescription of Uncinaria lucasi<br />

Stiles, a hookworm of seals. <strong>Parasitology</strong> 38:160-<br />

162.<br />

Dailey, M. D., and B. L. Hill. 1970. A survey of<br />

metazoan parasites infecting the California (Zalophus<br />

californianus) and Steller (Eumetopias jubatus)<br />

sea lion. Bulletin of the Southern California<br />

Academy of Science 69:126-132.<br />

George-Nascimento, M., M. Lima, and E. Ortiz.<br />

1992. A case of parasite-mediated competition?<br />

Phenotypic differentiation among hookworms Uncinaria<br />

sp. (Nematoda: Ancylostomatidae) in<br />

sympatric and allopatric populations of South<br />

American sea lions Otaria byronia, and fur seals<br />

Arctocephalus australis (Carnivora: Otariidae).<br />

Marine Biology 112:527-533.<br />

Kolevatova, A. I., le. la. Serebriannikov, V. V.<br />

Fornin, and L. A. Safronova. 1978. Uncinaria<br />

lucasi in fur seals. Prophylaxis and treatment of<br />

agricultural animal diseases. Trudy Kirovskogo<br />

Sel'skokhoziaistvennogo Instituta 61:49-54. (In<br />

Russian.)<br />

Lucas, F. A. 1899. The causes of mortality among seal<br />

pups. Pages 75—98, Plates 16—21 in The Fur Seals<br />

and Fur Seal Islands of the North Pacific Ocean<br />

(David Starr Jordan Report, 1899) Part 3. Washington,<br />

D.C.<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Lyons, E. T. 1963. Biology of the hookworm, Uncinaria<br />

lucasi Stiles, 1901, in the northern fur seal,<br />

Callorhinus ursinus Linn, on the Pribilof Islands,<br />

Alaska. Ph.D. Dissertation, Colorado <strong>State</strong> University,<br />

Fort Collins. 87 pp., 5 pis.<br />

. 1994. Vertical transmission of nematodes: emphasis<br />

on Uncinaria lucasi in northern fur seals<br />

and Strongyloides westeri in equids. Journal of the<br />

Helminthological Society of Washington 61:169-<br />

178.<br />

, R. L. DeLong, S. R. Melin, and S. C. Tolliver.<br />

1997. Uncinariasis in northern fur seal and<br />

California sea lion pups from California. Journal<br />

of Wildlife Diseases 33:848-852.<br />

, J. H. Drudge, and S. C. Tolliver. 1976.<br />

Studies on the development and chemotherapy of<br />

larvae of Parascaris equorum (Nematoda:Ascaridoidea)<br />

in experimentally and naturally infected<br />

foals. Journal of <strong>Parasitology</strong> 62:453-459.<br />

, M. C. Keyes, and J. Conlogue. 1978. Activities<br />

of dichlorvos or disophenol against the hookworm<br />

{Uncinaria lucasi) and sucking lice of<br />

northern fur seal pups (Callorhinus ursinus) on St.<br />

Paul Island, Alaska. Journal of Wildlife Diseases<br />

14:455-464.<br />

, and O. W. Olsen. 1960. Report on the seventh<br />

summer of investigations on hookworms,<br />

Uncinaria lucasi Stiles, 1901, and hookworm diseases<br />

of fur seals, Callorhinus ursinus Linn, on<br />

the Pribilof Islands, Alaska, from 15 June to 3<br />

October 1960. U. S. Department of the Interior,<br />

Fish and Wildlife Service, Washington, D.C. 26<br />

pp.<br />

Mizuno, A. 1997. Ecological study on the hookworm,<br />

Uncinaria lucasi, of northern fur seal, Callorhinus<br />

ursinus, in Bering Island, Russia. Japanese Journal<br />

of Veterinary Research 45:109-110. (English<br />

summary of a thesis presented to the School of<br />

Veterinary Medicine, Hokkaido University, Japan.)<br />

Olsen, O. W. 1954. Report on the third summer of<br />

investigations on hookworms, Uncinaria lucasi<br />

Stiles, 1901, and hookworm disease of fur seals,<br />

Callorhinus ursinus Linn., on the Pribilof Islands,<br />

Alaska, from May 21 to September 17, 1953. U.S.<br />

Department of the Interior, Fish and Wildlife Service,<br />

Washington, D.C. 117 pp.<br />

. 1956. Report on the fifth summer of investigations<br />

on the hookworm, Uncinaria lucasi Stiles,<br />

1901 and hookworm disease in fur seals, Callorhinus<br />

ursinus Linn., on the Pribilof Islands, Alaska,<br />

from June 9 to August 20, 1955. U.S. Department<br />

of the Interior, Fish and Wildlife Service,<br />

Washington, D.C. 81 pp.<br />

. 1958. Hookworms, Uncinaria lucasi Stiles,<br />

1901, in fur seals, Callorhinus ursinus (Linn.), on<br />

the Pribilof Islands. Pages 152-175 in Transactions<br />

of Twenty-third North American Wildlife<br />

Conference, Wildlife Management Institute,<br />

Washington, D.C.<br />

, and E. T. Lyons. 1962. Life cycle of the<br />

hookworm, Uncinaria lucasi Stiles, of northern<br />

fur seals, Callorhinus ursinus on the Pribilof Is-


lands in the Bering Sea. Journal of <strong>Parasitology</strong><br />

48 (supplement):42-43.<br />

-, and . 1965. Life cycle of Uncinaria<br />

lucasi Stiles, 1901 (Nematoda: Ancylostomatidae)<br />

of fur seals, Callorhinus ursinus Linn., on the<br />

Pribilof Islands, Alaska. Journal of <strong>Parasitology</strong><br />

51:689-700.<br />

Riley, F. 1961. Fur seal industry of the Pribilof Islands,<br />

1786-1960. Fishery Leaflet 516, U.S. Department<br />

of the Interior, Fish and Wildlife Service,<br />

Washington, D.C. 14 pp.<br />

Stewart, T. B., W. M. Stone, and O. G. Marti. 1976.<br />

Strongyloides ransomi: prenatal and transmammary<br />

infection of pigs of sequential litters from<br />

LYONS ET AL.—HOOKWORMS IN NORTHERN FUR SEAL PUPS 223<br />

Editors' Acknowledgments<br />

dams experimentally exposed as weanlings.<br />

American Journal of Veterinary Research 37:541-<br />

544.<br />

Stiles, C. W. 1901. Uncinariosis (Anchylostomiasis)<br />

in man and animals in the United <strong>State</strong>s. Texas<br />

Medical News 10:523-532.<br />

Trites, A. W. 1992. Northern fur seals: why have they<br />

declined? Aquatic Mammals 18:3-18.<br />

York, A. E., R. G. Towell, R. R. Ream, J. D. Baker,<br />

and B. W. Robson. <strong>2000</strong>. Population assessment,<br />

Pribilof Islands, Alaska. Pages 7-26 in B. W. Robson,<br />

ed. Fur Seal Investigations, 1998. U.S. Department<br />

of Commerce, NOAA Technical Memorandum<br />

NMFS-AFSC-113. 101 pp.<br />

We would like to acknowledge, with thanks, the following persons for providing their valuable<br />

help and insights in reviewing manuscripts for the Journal of the Helminthological Society of<br />

Washington and <strong>Comparative</strong> <strong>Parasitology</strong>: Omar Amin, Roy Anderson, Hisao Arai, Ann Barse,<br />

Diane Barton, Jeffrey Bier, Walter Boeger, Matthew Bolek, Daniel Brooks, Richard Buckner, Charles<br />

Bursey, Albert Bush, Rebecca Cole, Gary Conboy, Murray Dailey, Raymond Damian, David Dean,<br />

Donald Duszynski, Barbara Doughty, Tommy Dunagan, Larry Duncan, Ralph Eckerlin, William<br />

Font, Chris Gardner, Tim Goater, Stephen Goldberg, Richard Heckmann, Sherman Hendrix, Eric<br />

Hoberg, Jane Huffman, Renato Inserra, John Janovy, Hugh Jones, James Joy, Il-Hoi Kim, Michael<br />

Kinsella, Delane Kritsky, Ralph Lichtenfels, Donald Linzey, Austin Maclnnis, David Marcogliese,<br />

Lillian Mayberry, Chris McAllister, Robert Miller, Serge Morand, Michael Moser, Patrick Muzzall,<br />

Steve Nadler, Kazuya Nagasawa, Ronald Neafie, Paul Nollen, John Oaks, David Oetinger, Kazuo<br />

Ogawa, Niels 0rnbjerg, Wilbur Owen, Michael Patrick, Danny Pence, Gary Pfaffenberger, Oscar<br />

Pung, Dennis Richardson, Larry Roberts, Klaus Rohde, Wesley Shoop, Allen Shostak, Scott Snyder,<br />

Robert Sorensen, Jane Starling, Mauritz ("Skip") Sterner, Vernon Thatcher, Dennis Thoney, John<br />

Ubelaker.<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 224-229<br />

Life History of Spiroxys hanzaki Hasegawa, Miyata, et Doi, 1998<br />

(Nematoda: Gnathostomatidae)<br />

HIDEO HASEGAWA,1'3 TOSHIO Doi,2 AKIKO FunsAKi,1 AND AKIRA MIYATA'<br />

1 Department of Biology, Oita Medical University, Hasama, Oita 879-5593, Japan<br />

(e-mail: hasegawa@oita-med.ac.jp) and<br />

2 Suma Aqualife Park, Wakamiya, Suma, Kobe, Hyogo 654-0049, Japan<br />

ABSTRACT: The life history of Spiroxys hanzaki Hasegawa, Miyata, et Doi, 1998 (Nematoda: Gnathostomatidae),<br />

a stomach parasite of the Japanese giant salamander, Andrias japonicus (Temminck, 1836) (Caudata: Cryptobranchidae),<br />

was studied. The eggs developed in water to liberate sheathed second-stage larvae with a cephalic<br />

hook. They were ingested by the cyclopoid copepods, Mesocyclops dissimilis Defaye et Kawabata, 1993, and<br />

Macrocyclops albidus (Jurine, 1820) and developed to infective third-stage larvae in the hemocoel. Natural<br />

infections with third-stage larvae were also found in the cobitid loaches, Misgurnus anguillicaudatus (Cantor,<br />

1842) and Cobitis biwae (Jordan et Snyder, 1901). The largest third-stage larva from A. japonicus had almost<br />

the same body size as the smallest immature adult.<br />

KEY WORDS: Spiroxys hanzaki, Nematoda, Gnathostomatidae, life history, Andrias japonicus, Japanese giant<br />

salamander, Caudata, Cryptobranchidae, Copepoda, Mesocyclops, Macrocyclops, Japan.<br />

The Japanese giant salamander, Andrias japonicus<br />

(Temminck, 1836) (Cryptobranchidae),<br />

is an endangered amphibian distributed only in<br />

West Japan and protected by Japanese national<br />

law. From this salamander, a new nematode, Spiroxys<br />

hanzaki Hasegawa, Miyata, et Doi, 1998<br />

(Gnathostomatidae), was described recently<br />

(Hasegawa et al., 1998). Although it was suggested<br />

that the salamander acquired the infection<br />

by ingesting freshwater fish harboring the infective<br />

stage of S. hanzaki (Hasegawa et al., 1998),<br />

there is insufficient evidence for this. Recently,<br />

viable eggs of S. hanzaki were unexpectedly<br />

available, allowing attempts to experimentally<br />

infect copepods as intermediate hosts. In addition,<br />

freshwater fish captured in the rivers where<br />

the giant salamanders live were examined for<br />

larvae of S. hanzaki. The larval stages were also<br />

compared with those observed in the definitive<br />

host. We present herein the results of these observations,<br />

with a discussion on the developmental<br />

stages of gnathostomatoid nematodes.<br />

Materials and Methods<br />

Experiments on embryonic and larval<br />

development<br />

On 4 July 1998, 1 A. japonicus reared in the Suma<br />

Aqualife Park, Kobe, Hyogo Prefecture, Japan, vomited<br />

a half-digested loach, Misgurnus anguillicaudatus<br />

Cantor, 1842, that had been given on the previous day<br />

as food. Many individuals of S. hanzaki at various de-<br />

Corresponding author.<br />

224<br />

Copyright © 2011, The Helminthological Society of Washington<br />

velopmental stages were found invading the skin, muscles,<br />

and viscera of the loach. The loach was kept at<br />

4°C and transported to the Department of Biology, Oita<br />

Medical University, for further examination. On arrival<br />

(6 July 1998), all the worms were still alive. Eggs were<br />

obtained by tearing the uteri of 2 gravid females.<br />

Meanwhile, the remaining worms were fixed with 70%<br />

ethanol at 70°C for routine morphological examination<br />

or were stored at — 25°C for future biochemical analysis.<br />

The eggs were incubated in distilled water in a Petri<br />

dish (9 cm in diameter) at 15°C for 11 days, and then<br />

the temperature was raised to 20°C to facilitate hatching.<br />

When larvae hatched, 1 or 2 were transferred by<br />

a capillary pipette to each of several small Petri dishes<br />

(3 or 4 cm in diameter) containing about 5 ml of pond<br />

water. Copepods were collected in a nearby pond or<br />

paddy with a plankton net and were introduced to the<br />

dishes containing S. hanzaki larvae. Each copepod was<br />

observed daily thereafter under a stereomicroscope to<br />

examine the development of 5. hanzaki larvae inside<br />

the body. Identification of copepods was based on<br />

Ueda et al. (1996, 1997).<br />

Some newly hatched larvae were fixed by slight<br />

heating to observe their morphology. Infected copepods<br />

were dissected in physiological saline at various<br />

days of infection, and recovered larvae were killed by<br />

slight heating or by placing them in 70% ethanol at<br />

70°C. Heat-killed larvae were examined immediately,<br />

whereas those fixed in 70% ethanol were cleared in<br />

glycerol-alcohol solution by evaporating the alcohol,<br />

mounted on a glass slide with 50% glycerol aqueous<br />

solution, and observed under a Nikon Optiphot microscope<br />

equipped with a Nomarski differential interference<br />

apparatus. Measurements are in micrometers unless<br />

otherwise stated.<br />

Larvae parasitic in naturally infected fish<br />

Between May and November 1998, the following<br />

fish were netted in the Hatsuka River and the Okuyama


River, Kobe, Hyogo Prefecture, Japan, and were examined<br />

for larvae of Spiroxys: from the Hatsuka River:<br />

47 Zacco temmincki (Temminck et Schlegel, 1846)<br />

(Cyprinidae) (body length 37-137 mm); 2 Morokojouyi<br />

(Jordan et Snyder, 1901) (Cyprinidae) (54-58 mm);<br />

2 Pungutungia herzi Herzenstein, 1892 (Cyprinidae)<br />

(90-95 mm); 4 Misgurnus anguillicaudatus (Cantor,<br />

1842) (Cobitidae) (46-80 mm); 13 Cobitis biwae Jordan<br />

et Snyder, 1901 (Cobitidae) (38-95 mm); 4 Rhinogobius<br />

flumineus (Mizuno, 1960) (Gobiidae) (33-45<br />

mm); and 1 Odontobutis obscura (Temminck et Schlegel,<br />

1845-1846) (Gobiidae) (96 mm); from the Okuyama<br />

River: 10 Z. temmincki (41-75 mm). Their viscera<br />

were pressed between 2 thick glass plates and<br />

observed under a stereomicroscope with transillumination<br />

to detect Spiroxys larvae. The remaining portions<br />

of the fish were minced and digested with artificial<br />

gastric fluid for 3 hr at 37°C. The residues were<br />

transferred to a Petri dish and examined for nematode<br />

larvae under a stereomicroscope. Larvae detected were<br />

processed as described above for morphological observation.<br />

Scientific names of the fishes follow those<br />

adopted by Masuda et al. (1984).<br />

The third-stage larvae of Spiroxys japonica Morishita,<br />

1926, from the Asian pond loach, M. anguillicaudatus,<br />

and frogs, Rana nigromaculata Hallowell,<br />

1860, and Rana rugosa Schlegel, 1838, captured in<br />

Niigata and Akita Prefectures, northeastern Japan,<br />

where A. japonicus does not occur, were examined for<br />

comparison.<br />

Voucher nematode specimens were deposited in the<br />

United <strong>State</strong>s National Parasite Collection (USNPC),<br />

Beltsville, Maryland, U.S.A., Nos. 89629-89638.<br />

Results<br />

Embryonic development<br />

When the culture started, the nematode eggs<br />

contained 1- to 4-cell-stage embryos. After 2<br />

days of culture, they developed to 16-cell to<br />

morula stage. On days 7 and 8 of culture, tadpole-stage<br />

embryos were seen. On day 10, firststage<br />

larvae showed movement within the eggshell,<br />

and some larvae began to molt to become<br />

second stage. On day 11, molted larvae were<br />

observed. On day 18, eggs began to hatch (Fig.<br />

1), and hatched second-stage larvae were still<br />

enclosed in a sheath, adhered by the tips of their<br />

tails to the bottom of the culture dish. They seldom<br />

swam in the water.<br />

MORPHOLOGY OF HATCHED SECOND-STAGE LAR-<br />

VAE (n = 4): Stumpy worm with tapered posterior<br />

portion (Fig. 1). Enclosed within doublelayered<br />

sheath: outer layer lacking striations,<br />

and inner layer with reticular markings (Figs. 2,<br />

3). Length 330-435, maximum width 25-32.<br />

Anterior end with dorsal sclerotized hooklet<br />

with elongated base (Fig. 2). Esophagus 118-<br />

173 long, widened posteriorly and narrowed at<br />

HASEGAWA ET AL.—LIFE HISTORY OF SPIROXYS HANZAKI 225<br />

level of nerve ring. Nerve ring 65-85 from anterior<br />

extremity. Intestinal wall with brown granules.<br />

Excretory pore, genital primordium, and<br />

anus indiscernible.<br />

Development in intermediate host<br />

Several species of copepods were used for experimental<br />

infection. Preliminary trials revealed<br />

that Mesocyclops dissimilis Defaye et Kawabata,<br />

1993, Macrocyclops albidus (Jurine, 1820), and<br />

3 species of unidentified cyclopoids readily ingested<br />

the hatched larvae, but infection was established<br />

only in the former 2 species. The other<br />

species could not tolerate the infection and soon<br />

died. Hence, the following results were based on<br />

the experiments using M. dissimilis and M. albidus<br />

as intermediate hosts.<br />

After being ingested by the copepods, the larvae<br />

soon migrated to the hemocoel of the host<br />

(Fig. 4). The sheath was not observed in the larvae<br />

that had migrated to the hemocoel. Among<br />

31 M. dissimilis challenged, 15 were found to<br />

ingest the larvae, whereas worm uptake was not<br />

confirmed in the remaining individuals. The larvae<br />

disappeared from the hemocoel of 3 M. dissimilis<br />

by day 7 after infection. The copepods<br />

harboring S. hanzaki larvae became emaciated,<br />

5 of them died by day 10, and 6 more died by<br />

day 20. The larvae recovered by dissecting these<br />

dead copepods showed little development, still<br />

possessing the cephalic hooklet (Fig. 5). In 1 M.<br />

dissimilis, disseminated fatal infection with unidentified<br />

flagellates was caused after migration<br />

of S. hanzaki larvae. Ultimately, only 1 M. dissimilis<br />

survived for more than 25 days. When<br />

dissected on the 35th day of infection, this copepod<br />

harbored 1 living third-stage larva and 1<br />

dead second-stage larva.<br />

Among 10 M. albidus challenged, only 2 were<br />

found to harbor the larvae in the hemocoel on<br />

day 2 after infection, but 1 of them died by day<br />

10. The remaining individual died on day 24,<br />

but 1 third-stage larva was recovered from it by<br />

dissection. The control copepods, 36 M. dissimilis<br />

and 10 M. albidus, were not observed to<br />

be infected with any nematode throughout the<br />

experiment.<br />

MORPHOLOGY OF THE SECOND-STAGE LARVAE<br />

COLLECTED FROM THE INFECTED COPE-<br />

PODS: Identical with that of the hatched larvae<br />

but lacking sheaths; size gradually increased as<br />

the duration of infection lengthened. On day 8<br />

after infection, length 313-333, maximum width<br />

Copyright © 2011, The Helminthological Society of Washington


226 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Figures 1-3. Second-stage larva of Spiroxys hanzaki. 1. Hatching larva (scale bar = 50 urn). 2. Anterior<br />

portion, lateral view, showing cephalic hooklet (arrow) and double-layered cuticle (darts) (scale bar = 25<br />

(Am). 3. Inner layer of cuticle showing reticulated nature (darts) (scale bar = 25 jxm).<br />

Figure 4. Spiroxys hanzaki larva (arrow) in the hemocoel of Mesocyclops dissimilis on day 1 after<br />

exposure (scale bar = 200 u.m).<br />

Figure 5. Second-stage larva collected from the hemocoel of M. dissimilis at 15 days after infection.<br />

Arrow indicates cephalic hooklet (scale bar = 50 u.m).<br />

Figures 6-8. Third-stage larvae collected from the infected copepod, lateral view (scale bars = 50 u,m).<br />

6. Anterior portion. 7. Genital primordium (arrow). 8. Posterior portion.<br />

Figures 9-10. Third-stage larva of S. hanzaki in naturally infected sand loach, Cobitis biwae, lateral<br />

view (scale bars = 50 (xm). 9. Anterior portion. 10. Posterior portion.<br />

Copyright © 2011, The Helminthological Society of Washington


HASEGAWA ET AL.—LIFE HISTORY OF SPIROXYS HANZAKl 227<br />

Table 1. Measurements of third-stage larvae of Spiroxys hanzaki collected from experimentally-infected<br />

copepods, naturally infected fish, and salamanders (measurements in micrometers unless stated otherwise).<br />

No. measured<br />

Body length, mm<br />

Maximum width<br />

Nerve ringt<br />

Excretory poret<br />

Deiridst<br />

Esophagus length, mm<br />

Esophagus width<br />

Genital primordium, mm:|:<br />

Tail length<br />

* Advanced third-stage larvae.<br />

t Distance from cephalic extremity.<br />

+ Distance from caudal extremity.<br />

Mesocyclops<br />

dissimilis and<br />

Macrocyclops<br />

albidus<br />

2<br />

1.39-1.80<br />

61-80<br />

140-198<br />

175-219<br />

220-296<br />

0.50-0.70<br />

32-34<br />

0.46-0.47<br />

45-56<br />

18—19 at posterior esophagus level, nerve ring<br />

69—72 from anterior extremity and esophagus<br />

115-143 long (n = 2). On day 15, length 345,<br />

maximum width 18, nerve ring 75 from anterior<br />

extremity and esophagus 148 long (n =1).<br />

MORPHOLOGY OF THE THIRD-STAGE LARVAE<br />

COLLECTED FROM THE INFECTED COPEPODS: Body<br />

slender. Cuticle with fine transverse striations.<br />

Lateral alae absent. Anterior extremity with lateral<br />

pseudolabia with trilobed internal sclerotized<br />

structure of which dorsal and ventral lobes<br />

much smaller than median lobe, directed anteriorly<br />

(Figs. 6, 17). Large submedian papillae<br />

and amphidial pore present (Fig. 17). Esophagus<br />

club-shaped. Intestinal Wall densely packed with<br />

brown granules. Genital primordium with elongated<br />

2 branches extending anteriorly and posteriorly<br />

(Fig. 7). Tail conical, with prominent<br />

phasmidial pores and blunt extremity (Fig. 8).<br />

Measurements are presented in Table 1.<br />

Natural infection of fish with Spiroxys larva<br />

A total of 83 individuals of 7 fish species belonging<br />

to 3 families was examined during the<br />

Misgtirnus<br />

anguillicaudatus and<br />

Cobitis biwae<br />

3<br />

1.33-2.01<br />

46-56<br />

118-144<br />

149-205<br />

226-304<br />

0.41-0.58<br />

28-38<br />

0.43-0.77<br />

56-69<br />

Andrias<br />

japonicus<br />

2<br />

1.76-2.00<br />

56-90<br />

176-143<br />

214-190<br />

293-296<br />

0.57-0.65<br />

28-32<br />

0.54-0.65<br />

54-64<br />

Andrias<br />

japonicus<br />

5~'~<br />

6.70-9.00<br />

208-286<br />

384-455<br />

462-539<br />

666-813<br />

1.83-2.16<br />

78-102<br />

2.36-3.02<br />

150-183<br />

period from May to November 1998. Only 1 M.<br />

anguillicaudatus and 2 sand loach, Cobitis biwae<br />

(Jordan et Snyder, 1901), were found to be<br />

infected each with 1 larva of Spiroxys. Two of<br />

the larvae were found encysted on the stomach<br />

wall and liver surface, whereas the remaining<br />

larva was recovered by artificial digestion. The<br />

morphology was identical with that of the larvae<br />

recovered from the experimentally infected copepods<br />

(Figs. 9, 10, 18). Measurements are also<br />

comparable with those of the third-stage larvae<br />

from the experimentally infected copepods as<br />

shown in Table 1.<br />

The third-stage larva of S. hanzaki is readily<br />

distinguished from that of S. japonica, because<br />

the latter has inwardly curved dorsal and ventral<br />

lobes of the internal sclerotized structure in the<br />

pseudolabium (Figs. 11, 19).<br />

Morphology of 5. hanzaki larvae and<br />

immature adults vomited from A. japonicus<br />

THIRD-STAGE LARVAE (Figs. 12-14): Morphology<br />

comparable with those from the exper-<br />

Figure 11. Third-stage larva of Spiroxys japonica collected from pond loach, Misgurnus anguillicaudatus<br />

from Hachiro-gata, Akita Prefecture, Japan, lateral view, showing inwardly bent dorsal and ventral<br />

lobes of the sclerotized structure in pseudolabium (scale bar = 50 |xm).<br />

Figures 12-14. Smallest third-stage larva of S. hanzaki collected from Andrias japonicus, lateral view<br />

(scale bars = 50 jxm). 12. Anterior portion. 13. Genital primordium (arrow). 14. Posterior portion.<br />

Figures 15, 16. Advanced third-stage larva vomited by A. japonicus, lateral view (scale bars = 50 |xm).<br />

15. Anterior extremity. 16. Posterior extremity.<br />

Copyright © 2011, The Helminthological Society of Washington


228 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Figures 17-19. Cephalic extremities of third-stage larvae of Spiroxys spp., lateral view (scale bars =<br />

25 (Jim). 17, 18. Spiroxys hanzaki collected from experimentally infected copepod, Mesocyclops dissimilis<br />

(17), and naturally infected sand loach, Cobitis biwae (18). 19. Spiroxys japonica collected from naturally<br />

infected Rana rugosa (tadpole) in Akita, Akita Prefecture, Japan.<br />

imentally infected copepods or naturally infected<br />

fish. Measurements are stated in Table 1.<br />

ADVANCED THIRD-STAGE LARVAE (Figs. 15, 16):<br />

Morphology identical with that of the third-stage<br />

larvae described above but with much larger<br />

body (Table 1).<br />

IMMATURE ADULTS: Morphology identical<br />

with that of mature adults described in Hasegawa<br />

et al. (1998), but much smaller in size: males<br />

10.2-13.3 mm long (« = 5) and females 10.2-<br />

15.5 mm long (n = 5).<br />

Discussion<br />

Although only a few third-stage larvae were<br />

recovered from the experimentally infected copepods,<br />

it is apparent that 5. hanzaki utilizes copepods<br />

as its intermediate host, like most gnathostomatoids<br />

for which life histories have been<br />

elucidated (cf. Anderson, 1992). Compared with<br />

Spiroxys contortus (Rudolphi, 1819) and S. japonica<br />

(cf. Hedrick, 1935; Hasegawa and Otsuru,<br />

1978), S. hanzaki shows some different<br />

features in the life history. Hatched larvae are<br />

much larger (175-294 long in S. contortus, and<br />

148-207 long in S. japonica). The hatched larva<br />

attaches at the bottom, like that of 5. contortus,<br />

whereas the larva of S. japonica often swims in<br />

the water. Moreover, the period to attain the third<br />

stage in the copepods is much longer (10 to 14<br />

days and 6 to 8 days in S. contortus and S. japonica,<br />

respectively).<br />

The large size of the hatched larva and the<br />

slow development may be responsible for the<br />

high mortality rate of the infected copepods.<br />

Penetration of such a large larva through the alimentary<br />

canal wall of the copepod may result<br />

in perforation, through which pathogenic organisms<br />

could easily invade the hemocoel, as shown<br />

by the disseminated infection with flagellates in<br />

Copyright © 2011, The Helminthological Society of Washington<br />

the present experiment. Meanwhile, it is also<br />

probable that the larva in the hemocoel would<br />

stimulate some defense mechanism of the copepods<br />

to eliminate the invader, because the larvae<br />

often died without further development.<br />

The worm size and morphology of the thirdstage<br />

larvae from the copepods are similar to<br />

those of the smallest third-stage larva found in<br />

the salamander. This suggests that the third-stage<br />

larva developed in copepods could be infective<br />

to the final host. However, most of the gnathostomatoids<br />

require the second intermediate or<br />

paratenic hosts, often fish, in which the thirdstage<br />

larva becomes the so-called advanced third<br />

stage, that shows significant gain in body size<br />

but without essential morphological change (cf.<br />

Anderson, 1992). A similar pattern was also<br />

postulated for S. hanzaki (Hasegawa et al.,<br />

1998). Because the low tolerance of the copepods<br />

to the infection in our experiments prevented<br />

experimental infection of fish with raised<br />

larva, it remains unknown whether the larvae<br />

grow to the advanced third stage in fish.<br />

In most parasitic nematodes, the fourth stage<br />

exists between the infective third stage and fifth<br />

(adult) stage. However, the presence of the<br />

fourth-stage larva in Spiroxys is doubtful. Hedrick<br />

(1935) stated that the third and fourth molts<br />

of S. contortus were observed in the definitive<br />

host turtles but did not present the morphology<br />

of the fourth stage. In the life history study of<br />

S. japonica, Hasegawa and Otsuru (1978) could<br />

not find any larva that was distinguishable morphologically<br />

from the third-stage larva in the experimentally<br />

infected definitive host, frogs. Berry<br />

(1985) described the larvae of Spiroxys chelodinae<br />

Berry, 1985, collected from the stomach<br />

ulcers of Australian chelonians, as fourth stage,<br />

but the morphology resembles that of third-stage


larvae of 5. contortus or S. japonica. In the present<br />

study, the largest third-stage larva from the<br />

salamander had nearly the same body size as<br />

that of the smallest immature adult. These facts<br />

suggest that Spiroxys lacks a fourth larval stage.<br />

The presence of the fourth larval stage is also<br />

unclear- for other gnathostomatoids. In Gnathostorna<br />

spp., there has been no description of the<br />

fourth-stage larva. In the life history study of<br />

Gnathostoma procyonis Chandler, 1942, Ash<br />

(1962) termed the larva of which a cross-section<br />

was presented as the fourth stage in the figure<br />

caption. However, he did not use this term in the<br />

text or describe a stage morphologically different<br />

from both the third stage and the adult stage.<br />

Moreover, we recently observed that Gnathostoma<br />

doloresi Tubangui, 1925, larvae in molting<br />

to the adult stage had the cuticle with typical<br />

arrangement of cephalic booklets of the third<br />

stage (specimens courtesy of Dr. J. Imai). Further<br />

careful study is required to determine<br />

whether gnathostomatoids molt only once in the<br />

definitive host.<br />

Acknowledgments<br />

Sincere thanks are rendered to Dr. J. Imai, Miyazaki<br />

Medical <strong>College</strong>, Dr. M. Koga, Kyushu<br />

University School of Medicine, and Dr. H. Akahane,<br />

Fukuoka University School of Medicine,<br />

for their kindness in providing invaluable information<br />

on the development of Gnathostoma spp.<br />

Thanks are also extended to Dr. T. Yoshino, University<br />

of the Ryukyus, for his kindness in verifying<br />

the scientific names of the fish. This study<br />

HASEGAWA ET AL.—LIFE HISTORY OF SPIROXYS HANZAKI 229<br />

was partly supported by the grant-in-aid from<br />

the Ministry of Education, Science and Culture,<br />

Japanese Government, No. 11640700.<br />

Literature Cited<br />

Anderson, R. C. 1992. Nematode Parasites of Vertebrates.<br />

Their Development and Transmission.<br />

C.A.B International, Wallingford, U.K. 578 pp.<br />

Ash, L. R. 1962. Migration and development of Gnathostoma<br />

procyonis Chandler, 1942, in mammalian<br />

hosts. Journal of <strong>Parasitology</strong> 48:306-313.<br />

Berry, G. N. 1985. A new species of the genus Spiroxys<br />

(Nematoda: Spiruroidea) from Australian<br />

chelonians of the genus Chelodina (Chelidae).<br />

Systematic <strong>Parasitology</strong> 7:59-68.<br />

Hasegawa, H., A. Miyata, and T. Doi. 1998. Spiroxys<br />

hanzaki n. sp. (Nematoda: Gnathostomatidae) collected<br />

from the giant salamander, Andrias japonicus<br />

(Caudata: Cryptobranchidae), in Japan. Journal<br />

of <strong>Parasitology</strong> 84:831-834.<br />

, and M. Otsuru. 1978. Notes on the life cycle<br />

of Spiroxys japonica Morishita, 1926 (Nematoda:<br />

Gnathostomatidae). Japanese Journal of <strong>Parasitology</strong><br />

27:113-122.<br />

Hedrick, L. R. 1935. The life history and morphology<br />

of Spiroxys contortus (Rudolphi); Nematoda: Spiruridae).<br />

Transactions of the American Microscopical<br />

Society 54:307-335.<br />

Masuda, H., K. Amaoka, C. Araga, T. Uyeno, and<br />

T. Yoshino, eds. 1984. The Fishes of the Japanese<br />

Archipelago. Tokai University Press, Tokyo. 456<br />

pp.<br />

Ueda, H., T. Ishida, and J. Imai. 1996. Planktonic<br />

cyclopoid copepods from small ponds in Kyushu,<br />

Japan. I. Subfamily Eucyclopinae with descriptions<br />

of micro-characters on appendages. Hydrobiologia<br />

333:5=56.<br />

, , and . 1997. Planktonic cyclopoid<br />

copepods from small ponds in Kyushu, Japan.<br />

II. Subfamily Cyclopinae. Hydrobiologia<br />

356:61-71.<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 230-235<br />

Inducible Nitric Oxide Synthase in the Muscles of Trichinella sp.-<br />

Infected Mice Treated with Glucocorticoid Methylprednisolone<br />

KRYSTYNA BOCZON' AND BARBARA WARGIN<br />

Department of Biology and Medical <strong>Parasitology</strong>, Karol Marcinkowski University of Medical Sciences,<br />

61-701 Poznari, Poland (e-mail: kboczon@eucalyptus.usoms.poznan.pl)<br />

ABSTRACT: The dynamics of inducible nitric oxide synthase (iNOS) activity in mice infected with Trichinella<br />

spiralis larvae were followed between the first and tenth week postinfection (p.L). During infection with T.<br />

spiralis, a bimodal stimulation of iNOS activity to 371% of the control value by day 21 p.i. and to 285% by<br />

day 70 p.i. was observed. The first increase in iNOS activity was abolished by glucocorticoid treatment. In T.<br />

pseudospiralis infection, the dynamics of iNOS stimulation differed from that in mice infected with T. spiralis:<br />

a constant but much weaker stimulation of iNOS starting on day 21 p.i. lasted until the end of the study. The<br />

results suggest that nitric oxide synthase activity is induced in muscle of the mouse during trichinellosis and<br />

that nitric oxide may participate in the host's biochemical defense mechanism.<br />

KEY WORDS: iNOS, inducible nitric oxide, Trichinella spiralis, Trichinella pseudospiralis, muscle, mouse,<br />

glucocorticoid treatment, methylprednisolone.<br />

The past decade has witnessed an increase in<br />

the number of papers devoted to the role of nitric<br />

oxide (NO) synthase in the pathogenesis of<br />

many diseases. Part of this surge in interest is<br />

related to the discovery of a role in both signal<br />

transduction and cell toxicity for NO. Induction<br />

of inducible nitric oxide synthase (iNOS) has<br />

been observed in the course of many human diseases.<br />

The parasitic infections investigated until<br />

now include malaria (Tsuji et al., 1995); leishmaniasis<br />

(Stenger et al., 1996); and toxoplasmosis<br />

(Holscher et al., 1998). The role of NO in<br />

killing protozoans of the genus Leishmania was<br />

studied in greater detail as early as 1993 (Callahan<br />

et al., 1993), when it was established that<br />

the course of the disease is dependent to a considerable<br />

extent on the type of lymphokines generated<br />

by T lymphocytes. During infection with<br />

such protozoans as Trypanosoma cruzi Chagas,<br />

1909 (Rottenberg et al., 1996), or Toxoplasma<br />

gondii Nicolle et Manceaux, 1908 (Hayashi et<br />

al., 1996), NO has both antiparasitic and immunosuppressive<br />

effects. Recent publications<br />

have also reported modulation of the expression<br />

of messenger RNA responsible for tumor necrosis<br />

factor—and prostaglandin E2-independent<br />

synthesis of iNOS and production of NO in Entamoeba<br />

histolytica Schaudinn, 1903, infection<br />

(Wang et al., 1994). The type of free radicals<br />

contributing to pathogenesis in specific parasitic<br />

invasion depends on the developmental stage of<br />

1 Corresponding author.<br />

230<br />

Copyright © 2011, The Helminthological Society of Washington<br />

the parasite, and the protective function of NO<br />

seems to be tissue-specific (Scharton-Kersten et<br />

al., 1997).<br />

Nitric oxide generated by nitrogen free radicals<br />

(RNI), specifically one generated in inflammatory<br />

conditions by the inducible form of NOS<br />

(iNOS), is associated with macrophages and<br />

plays a fundamental role in killing or suppressing<br />

various pathogens (Gross and Wolin, 1995).<br />

The mechanism whereby NO influences the cell<br />

includes, among others, an effect on both respiration<br />

and oxygen potential in mitochondria<br />

and Fe-S proteins engaged in the Krebs cycle<br />

and in electron transport (Kroncke et al., 1995).<br />

In the case of NO overproduction, the concentration<br />

of oxygen in the environment plays an<br />

important role in regulating the functions of mitochondria.<br />

The balance between RNI and oxygen<br />

free radicals (ROS) is of special importance.<br />

Nitric oxide also participates in modulating<br />

enteritis during the intestinal phase of infection<br />

with Trichinella spiralis Owen, 1935, since inflammatory<br />

changes in the intestine of animals<br />

infected with T. spiralis were eliminated with a<br />

specific iNOS inhibitor. This suggests that iNOS<br />

may participate in the disease process associated<br />

with intestinal invasion by adult forms of T.<br />

spiralis (Hogaboam et al., 1996) and may<br />

through its influence on enteritis play an important<br />

role in rejection of adult worms.<br />

Our laboratory proposed a hypothesis that<br />

RNI may also play a role in protective mechanisms<br />

during the muscular phase of trichinellos-


is. Using histochemical methods our group demonstrated<br />

NOS in basophilically transformed<br />

muscle fibers in T. spiralis—infected mice (Hadas<br />

et al., 1999). In a separate paper we reported on<br />

the participation of ROS in the biochemical protective<br />

mechanisms in host muscle infected with<br />

T. spiralis larvae (Wandurska-Nowak et al.,<br />

1998). In the same paper we demonstrated that<br />

administration of the glucocorticoid methylprednisolone<br />

had a profound effect on the activity of<br />

antioxidant enzymes that were examined (superoxide<br />

dismutase [SOD] and peroxidase). According<br />

to Connors and Moncada (1991), glucocorticoid<br />

also inhibits iNOS.<br />

The initiation of research on the participation<br />

of iNOS in biochemical defense mechanisms of<br />

the host in T. spiralis infection was also important<br />

from the point of view of its possible participation<br />

in the mechanism of uncoupling of oxidative<br />

phosphorylation, which can be observed<br />

in the mitochondria of tissue infected with helminths<br />

(Michejda and Boczori, 1972; Van den<br />

Bosche et al., 1980; Boczori and Bier, 1986;<br />

Ruble et al., 1989). It was shown that the expected<br />

temporal correlation between the increase<br />

in the activity of SOD and peroxidase and the<br />

peaks in trichinellosis phosphorylation uncoupling<br />

did not occur (Wandurska-Nowak et al.,<br />

1998).<br />

The objectives of the present investigation<br />

were to determine 1) quantitative changes in the<br />

activity of iNOS in muscles from hosts infected<br />

with T. spiralis or Trichinella pseudospiralis<br />

Garkavi, 1976, and 2) if glucocorticoid prevents<br />

changes in the quantity of NO generated in infected<br />

tissues.<br />

Materials and Methods<br />

Experimental tissue consisted of muscles removed<br />

from uninfected mice (2-mo-old female mice, strain<br />

BALB/C) and from mice infected per os with 700-800<br />

infective larvae of either T. spiralis (strain MSUS/PO/<br />

60/ISS3) or T. pseudospiralis (strain MPRO/US/72/<br />

ISS13). The infective larvae obtained after pepsin-HCl<br />

digestion after about 2 hr for T. spiralis larvae and<br />

about 1-1.5 hr for T. pseudospiralis were administered<br />

per os to mice anesthetized with ether. The mice were<br />

killed by decapitation. The amount of larvae per 1 g<br />

of muscle tissue obtained after pepsin-HCl digestion<br />

at 6-8 wk post-infection (p.i.) were 10,000-12,000 and<br />

5,000 for T. spiralis and T. pseudospiralis, respectiveiy.<br />

Mice were bred and housed in the animal laboratory,<br />

which ensured approximately constant temperature,<br />

humidity, and ad libitum access to LMS Labofeed B<br />

BOCZON AND WARGIN—iNOS IN MOUSE MUSCLE 231<br />

(Feed and Concentrates Production Plant) granulated<br />

food and water.<br />

Only 1 group of animals infected with T. spiralis<br />

larvae was treated with methylprednisolone (Depomedrol<br />

[Jelfa, Poland], a drug with prolonged action)<br />

administered on day 7 p.i. by subcutaneous injection<br />

at a dose of 20 mg/kg of body weight. Quadriceps<br />

muscles from hind legs were removed and homogenized<br />

for 15 to 30 sec in a sucrose medium of the<br />

following content (in final concentration): 0.25 M sucrose,<br />

0.002 M EGTA, 0.01 M Tris HC1 buffer (pH<br />

7.3), and 20 jxl heparin with a concentration of 500<br />

units/g per 10 ml medium. The homogenate was centrifuged<br />

for 10 min at 4,500 rpm, and the resulting<br />

supernatant was centrifuged for 12 min at 15,000 rpm.<br />

The activity of iNOS was measured in the latter supernatant<br />

spectrophotometrically by Green's method as<br />

modified by Lepoivre (Lepoivre et al., 1989), using the<br />

following solutions: A) Griess' reagent containing<br />

0.5% sulphanilamide dissolved in 1 N HC1 and 0.15%<br />

/V-(l-napthyl) ethylendiamine mixed in a ratio of 1:1<br />

and B) consisting of (in final concentrations) 40 mM<br />

Tris HC1 buffer (pH 8.0), 2 mM NADPH, and 7 mM<br />

arginine. Enzyme activity was measured in 140 JJL! of<br />

supernatant after 30 min of incubation (to induce the<br />

enzyme activity) at 1-wk intervals at a wavelength of<br />

X = 540 nm in a cuvette containing 1,200 u,l of solution<br />

A and 100 u.1 of solution B. In some pilot experiments<br />

1.5 mM CaCl2 was added. Absorption readings<br />

were taken after a 30-min incubation period at a<br />

temperature of 24 °C, and NO concentration was determined<br />

using a NaNO2 standard curve. Protein was<br />

measured applying Lowry's method (Lowry et al.,<br />

1951).<br />

The measurements were carried out in 4 groups of<br />

animals: for T. spiralis-infected mice (I+NaCl), T.<br />

spiral is-infected mice under treatment (I+D), and also<br />

for 2 control groups (C+NaCl and C+D). Both infected<br />

and untreated mice (I+NaCl) and those from<br />

the respective control group (C + NaCl) were given intramuscular<br />

injections of 0.9% NaCl. Activity measured<br />

in the respective control groups was taken as<br />

100% for the calculation of percentage changes in such<br />

activity during T. spiralis infection and treatment.<br />

Analysis of variance or the Mann—Whitney test was<br />

used for statistical comparison between groups; P <<br />

0.01 (very significant) or


232 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Table 1. Activity of iNOS (in nmoles/mg of protein/min) in T. spiralis-infected (I+NaCl) and infected +<br />

methylprednisolone-treated mouse muscles (I+D).<br />

Days<br />

postinfection<br />

(d.p.i.)<br />

Controls<br />

7<br />

14<br />

21<br />

34<br />

42<br />

49<br />

70<br />

90<br />

Activity (±SEM)<br />

0.14 ± 0.01 (8)<br />

0.08 ± 0.007 (2)<br />

(P < 0.05)<br />

0.15 ± 0.007 (4)<br />

0.52 ± 0.1 (5)<br />

(P < 0.01)<br />

0. 1 1 ± 0.003 (6)<br />

0.07 ± 0.003 (5)<br />

(P < 0.05)<br />

0.07 ± 0.003 (6)<br />

(P < 0.05)<br />

0.40 ± 0.063 (6)<br />

(P < 0.01)<br />

0.04 ± 0.003 (3)<br />

(P < 0.05)<br />

I + NaCl*<br />

% of control<br />

57<br />

107<br />

371<br />

79<br />

50<br />

50<br />

285<br />

29<br />

I + D|<br />

Activity (±SEM)<br />

0.11 ± 0.003 (3)<br />

0.20 ± 0.026 (4)<br />

0.11 ± 0.053 (5)<br />

0.16 ± 0.003 (3)<br />

(P < 0.05)<br />

0.14 ± 0.056 (11)<br />

0.06 ± 0.003 (5)<br />

0.07 ± 0.003 (6)<br />

(P < 0.05)<br />

0.37 ± 0.254 (6)<br />

0.04 ± 0.001 (3)<br />

% of control<br />

* The statistically significant differences in column I+NaCl when the values of activity in infected mice were compared with<br />

normal mice. The number in parentheses is the number of measurements.<br />

t The comparison of the results from infected and infected and treated by glucocorticoid methylprednisolone mice was carried<br />

out using a Mann-Whitney test. The number in parentheses is the number of measurements.<br />

infected and treated, between 7 and 70 days p.i.,<br />

are presented in Table 1.<br />

The investigations of iNOS activity in muscles<br />

of mice infected with T. spiralis larvae<br />

showed a 2-stage increase in enzyme activity<br />

during the course of trichinellosis. An initial<br />

peak of activity was seen at 21 days p.i., and a<br />

second rise in activity occurred at 70 days pi<br />

(285% of the activity in the control group). Statistically<br />

the changes in activity during the stages<br />

of trichinellosis mentioned above varied significantly<br />

from controls (P < 0.01 or P < 0.05).<br />

The investigation of muscle enzyme activity<br />

was carried out on mice infected with T. spiralis<br />

and treated simultaneously with glucocorticoid<br />

methylprednisolone (I+D) at the same intervals<br />

as those for the group of infected and untreated<br />

animals. The I+D group mice had higher enzyme<br />

activity than those of the I+NaCl group<br />

on day 7 p.i. (182% of the control values) and<br />

at 70 days p.i. (up to 336% of the control value)<br />

with a statistically significant result compared<br />

with the control group (P < 0.01). At 90 days<br />

p.i. enzyme activity fell in a manner similar to<br />

that seen in T. spiralis-infected and untreated<br />

animals. Therefore, it may be assumed that<br />

methylprednisolone exerted a normalizing influence<br />

on iNOS activity only in the initial stage<br />

of the muscular phase, i.e., at day 21 p.i.<br />

Copyright © 2011, The Helminthological Society of Washington<br />

182<br />

100<br />

145<br />

127<br />

56<br />

64<br />

336<br />

The comparison of changes in iNOS activity<br />

in 2 infections, 1 caused by nonencysting T.<br />

pseudospiralis larvae and the other by encysting<br />

T. spiralis larvae, as presented in Figure 1, revealed<br />

a totally different dynamic for iNOS<br />

changes in muscles of infected mice. A statistically<br />

significant difference existed between the<br />

activity of the enzyme examined in mice infected<br />

with the T. spiralis and T. pseudospiralis invasion<br />

during all phases of trichinellosis (P <<br />

0.01). In general, unlike T. spiralis, infection<br />

with T. pseudospiralis was characterized by an<br />

absence of iNOS stimulation during the first<br />

weeks, while considerable stimulation (up to approximately<br />

280% of the control) lasted<br />

throughout and continued up to the end of the<br />

muscle phase (from 42 to 70 days p.i.).<br />

Discussion<br />

The biochemical defense mechanisms for killing<br />

T. spiralis larvae by eosinophils are mediated<br />

by peroxidase (POX), with inclusion of the<br />

process in which both neutrophils and eosinophils<br />

produce hypochlorous acid, which is toxic<br />

to the larvae of the parasite (Buys et al., 1981).<br />

The quantitative results of research on the activity<br />

of the inducible NOS isoform (iNOS) presented<br />

in this paper clearly indicate that this enzyme,<br />

supplying NO lethal to many parasites, is<br />

37


1.2 n<br />

1.0-<br />

|T. spiralis<br />

|T. pseudospiralis<br />

BOCZON AND WARGIN—iNOS IN MOUSE MUSCLE 233<br />

49 70 dpi<br />

Figure. 1. Comparison of the changes of the iNOS activity (mean ± SEM) during T. spiralis and T.<br />

pseudospiralis infection between 0-70 days postinfection. Values with * or ** (P


234 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY 2(X)0<br />

lymphocytes and suppression of NO production.<br />

In the present study, the immunosuppressive<br />

drug had no effect on iNOS induction in T. spira/zs-infected<br />

mice muscles. On the other hand,<br />

during the muscular phase of trichinellosis,<br />

methylprednisolone exerts a suppressive influence<br />

on the subpopulation of CDS lymphocytes<br />

(Boczori et al., unpublished data) with the degree<br />

of change dependent on the doses of larvae<br />

used to infect the host.<br />

Similar investigations on T. pseudospiralis<br />

were not performed because human cases of T.<br />

pseudospiralis have not been reported. Nevertheless,<br />

in severe cases of human trichinellosis<br />

caused by T. spiralis, glucocorticoid chemotherapy<br />

is popular, although still disputable.<br />

It is well known that larvae of T. pseudospiralis<br />

are less immunogenic than those of T. spiralis<br />

(Flockhart, 1986). Irrespective of the fact<br />

that they share 60% of antigens with T. spiralis<br />

larvae, they still possess 8 different proteins. In<br />

the intestinal phase, the adult form of T. pseudospiralis<br />

causes a much weaker inflammatory<br />

reaction than that of T. spiralis (Flockhart,<br />

1986). In the late muscular phase (70 days p.i-X<br />

the presence of T. pseudospiralis larvae resulted<br />

in a level of iNOS stimulation similar to that<br />

caused by T. spiralis larvae (approximately<br />

240%).<br />

Taking into account that the intensity of the<br />

T. pseudospiralis infection was 2 times lower<br />

than that of T. spiralis, we can assume that despite<br />

the lower immunogenicity of the former, it<br />

induced much stronger RNI generation. For example,<br />

on day 7 p.i. the iNOS activity recalculated<br />

per 1,000 muscle larvae of T. pseudospiralis/g<br />

of tissue was about 0.08 nmoles/mg of<br />

protein/min, and during the muscular phase<br />

measured on day 42 p.i. was 0.16 nmoles/mg of<br />

protein/min. In T. spiralis infection, the respective<br />

values for iNOS activity were 0.007 and<br />

0.006 nmoles/mg of protein/min, being about 10<br />

and 26 times lower, respectively, than in T. pseudospiralis<br />

infection.<br />

Thus, the continuous migration of T. pseudospiralis<br />

larvae causes a much greater degree<br />

of damage, but instead of an immunological response,<br />

a set of biochemical defense reactions,<br />

including RNI generation, take place.<br />

The agents present in excretory-secretory<br />

products of larvae induced iNOS for almost 3<br />

months. Still, in order to establish the possible<br />

duration of this effect, it would be necessary to<br />

Copyright © 2011, The Helminthological Society of Washington<br />

carry out a long-term study for at least 6 to 8<br />

mo p.i.<br />

Literature Cited<br />

Boczori, K., and J. W. Bier. 1986. Anisakis simplex:<br />

uncoupling of oxidative phosphorylation in the<br />

muscle of infected fish. Experimental <strong>Parasitology</strong><br />

61:270-279.<br />

, E. Hadas, E. Wandurska-Nowak, and M.<br />

A. Derda. 1996. Stimulation of antioxidants in<br />

muscles of Trichinella spiralis infected rats. Acta<br />

Parasitologica 41:136-138.<br />

Brown, G. C. 1995. Nitric oxide regulates mitochondrial<br />

respiration and cell functions by inhibiting<br />

cytochrome oxidase. Federation of European Biochemical<br />

Societies Letters 369:136-139.<br />

Butterworth, A. E. 1980. Eosinophils and immunity<br />

to parasites. Transactions of the Royal Society of<br />

Tropical Medicine and Hygiene 74 (supplement):<br />

38-43.<br />

Buys, J., R. Wever, R. Van Stigt, and E. J. Ruitenberg.<br />

1981. The killing of newborn larvae of<br />

Trichinella spiralis by eosinophils peroxidase in<br />

vitro. European Journal of Immunology 11:843—<br />

845.<br />

Callahan, H. L., E. R. James, and R. K. Crouch.<br />

1993. Oxidants and anti-oxidants in trypanosomiasis<br />

and leishmaniasis. Pages 81 — 110 in O. I.<br />

Aruoma, ed. Free Radicals in Tropical Diseases.<br />

Harwood Academic Publishers GmbH, Switzerland.<br />

Connors, K. J., and S. Moncada. 1991. Glucocorticoids<br />

inhibit the induction of nitric oxide synthase<br />

and the related cell damage in adenocarcinoma<br />

cells. Biochimica et Biophysica Acta 1097:227-<br />

231.<br />

Derda, M. 1998. Biochemical aspects of defence in<br />

experimental trichinellosis. Ph.D. Thesis, K. Marcinkowski<br />

University of Medical Sciences, Poznari,<br />

Poland. 51 pp.<br />

Despommier, D. D. 1990. The worm that would be<br />

virus. <strong>Parasitology</strong> Today 6:193—196.<br />

Flockhart, H. A. 1986. Trichinella speciation. <strong>Parasitology</strong><br />

Today 2:1-3.<br />

Gross, S. S., and M. S. Wolin. 1995. Nitric oxide<br />

pathophysiological mechanism. Pages 737—769 in<br />

J. F. Hoffman and P. De Weer, eds. Annual Review<br />

of Physiology 57. Annual Review, Inc., Palo Alto,<br />

California.<br />

Hadas, E., L. Gustowska, K. Boczori, and D. Janczewska.<br />

1999. Histochemical investigations of<br />

the biochemical defence mechanism in experimental<br />

trichinellosis. II. Nitric oxide synthase activity.<br />

Wiadomosci Parazytologiczne 45:5—15.<br />

Hayashi, S., C. C. Chan, R. Gazzinelli, and F. G.<br />

Roberge. 1996. Contribution of nitric oxide to the<br />

host parasite equilibrium in toxoplasmosis. Journal<br />

of Immunology 156:1476-1481.<br />

Hogaboam, C. M., S. M. Collins, and M. G. Blennerhassett.<br />

1996. Effects of oral L-NAME during<br />

Trichinella spiralis infection in rats. American<br />

Journal of Physiology 271 (2, part l):G338-346.<br />

Holscher, C., G. Kohler, U. Muller, H. Mossmann,


G. A. Schaub, and F. Brombacher. 1998. Defective<br />

nitric oxide effector functions lead to extreme<br />

susceptibility of Trypanosoma crwzz-infected<br />

mice deficient in gamma interferon receptor or<br />

inducible nitric oxide synthase. Infective Immunology<br />

66:1208-1215.<br />

Kroncke, K. D., K. Fehsel, and V. Kolb-Bachofen.<br />

1995. Inducible nitric oxide synthase and its product<br />

nitric oxide, a small molecule with complex<br />

biological activities. Biological Chemistry 376:<br />

327-343.<br />

Lepoivre, M., H. Boudbid, and J. F. Petit. 1989.<br />

Antiproliferative activity of gamma-interferon<br />

combined with lipopolysaccharide on murine adenocarcinoma:<br />

dependence on an L-arginine metabolism<br />

with production of nitrite and citrulline.<br />

Cancer Research 9:1970-1976.<br />

Lowry, O. H., W. J. Rosenbrough, A. Z. Faar, and<br />

R. J. Randall. 1951. Protein measurments with<br />

the Folin phenol reagent. Journal of Biological<br />

Chemistry 193:265-275.<br />

Michejda, J. W., and K. Boczori. 1972. Changes in<br />

bioenergetics of skeletal muscle mitochondria during<br />

experimental trichinosis in rats. Experimental<br />

<strong>Parasitology</strong> 31:161-171.<br />

Rottenberg, M. E., E. Castanos-Velez, R. Mesquita,<br />

O. G. Laguardia, P. Bibcrfeld, and A. 6rn.<br />

1996. Intracellular co-localization of T. cruzi and<br />

inducible nitric oxide synthase: Evidence for a<br />

dual pathway of iNOS induction. European Journal<br />

of Immunology 26:3203-3213.<br />

Ruble, C. J., C. A. Boehm, and F. L. Bygrave.1989.<br />

Aberrant energy-linked reactions in mitochondria<br />

isolated from the livers of rats infected with the<br />

liver fluke Fasciola hepatica. Biochemical Journal<br />

260:517-523.<br />

BOCZON AND WARGIN—iNOS IN MOUSE MUSCLE 235<br />

Scharton-Kersten, T. M., G. Yap, J. Magram, and<br />

A. Sher. 1997. Inducible nitric oxide is essential<br />

for host control of persistent but not acute infection<br />

with the intracellular pathogen Toxoplasma<br />

gondii. Journal of Experimental Medicine 185:<br />

1261-1273.<br />

Stenger, S., N. Donhauser, H. Thuring, M. Rollinghoff,<br />

and C. Bogdan. 1996. Reactivation of<br />

latent leishmaniasis by inhibition of inducible nitric<br />

oxide synthase. Journal of Experimental Medicine<br />

183:1501-1514.<br />

Tsuji, M., Y. Miyahira, R. S. Nussenzweig, M.<br />

Aguet, M. Reichel, and F. Zavala. 1995. Development<br />

of antimalarial immunity in mice lacking<br />

IFN-gamma receptor. Journal of Immunology 154:<br />

5338-5344.<br />

Wandurska-Nowak, E., K. Boczori, M. A. Derda, E.<br />

Hadas, and A. Matias-Guzek. 1998. The effect<br />

of methylprednisolone on the dynamics of oxygen<br />

uptake and peroxidase and superoxide dismutase<br />

activities in skeletal muscles of mice infected with<br />

Trichinella spimlis larvae. Acta Parasitologica43:<br />

54-56.<br />

Wang, W., K. Keller, and K. Chadee. 1994. Entamoeba<br />

histolytica modulates the nitric oxide synthase<br />

gene and nitric oxide production by macrophages<br />

for cytoxicity against amoebae and tumour<br />

cells. Immunology 83:601-610.<br />

Van den Bosche, H., H. Verhoeven, and H. Lauwers.<br />

1980. Uncoupling of liver mitochondria associated<br />

with fascioliasis in rats: normalization by<br />

closantel. Pages 699-704 in H. Van den Bosche,<br />

ed. The Host-Invader Interplay. Elsevier/North-<br />

Holland Biomedical Press, Amsterdam, New<br />

York, and Oxford.<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 236-240<br />

The Expulsion of Echinostoma trivolvis: Worm Kinetics and<br />

Intestinal Cytopathology in Jirds, Meriones unguiculatus<br />

TAKAHIRO FUJINO, K5 TOMONORI SHINOHARA,' KOICHI FuKUDA,2 HIDETAKA ICHIKAWA,S<br />

TOMOYUKI NAKANO,' AND BERNARD FRIED4<br />

1 Department of Biology, Faculty of Science, Yamagata University, 990-8560 Yamagata, Japan<br />

(e-mail: tfuji@sci.kj.yamagata-u.ac.jp),<br />

2 Center for Laboratory Animal Science, National Defense Medical <strong>College</strong>, Tokorozawa 359-8513, Japan<br />

(e-mail: kfukuda@cc.ndmc.ac.jp),<br />

3 Department of Medical Zoology, Kanazawa Medical University, Ishikawa 920-0293, Japan<br />

(e-mail: ichikawa@kanazawa-med.ac.jp), and<br />

4 Department of Biology, Lafayette <strong>College</strong>, Easton, Pennsylvania 18042, U.S.A.<br />

(e-mail: fried@lafvax.lafayette.edu)<br />

ABSTRACT: Worm kinetics and cytopathology of jirds, Meriones unguiculatus Milne-Edwards, 19<strong>67</strong>, infected<br />

with Echinostoma trivolvis (Cort, 1914) Kanev, 1985, were reported and compared with previous studies on<br />

echinostome infections in murine hosts. Seven jirds were each infected with 40 metacercarial cysts, and the<br />

worms were recovered at days 5, 8, 10, 12, 15, and 17 postinfection (p.i.). Worm recoveries were 35.4, 10.7,<br />

and 0.4 at days 5, 10, and 15 p.i., respectively. Worm expulsion occurred on about day 10 p.i., corresponding<br />

to the peak increase in the number of goblet cells at 24.3 ± 0.6/villus-crypt unit (VCU) at day 10 p.i. These<br />

data showed that worm expulsion of E. trivolvis in jirds occurred earlier than that in C3H and BALB/c mouse<br />

hosts. The difference in expulsion times and rates reflects differences in the peak number of goblet cells in the<br />

host intestines of jirds versus mice. The number of mucosal mast cells increased slightly and peaked at 1.1 ±<br />

0.32/10 VCU at day 10 in jirds. An increase in mucosal mast cells occurred earlier and was smaller in jirds<br />

than in BALB/c mice. Scanning electron microscope observations showed an irregular arrangement of microvilli<br />

in the small intestine of infected jirds. Transmission electron microscope observations also showed damage in<br />

the distal parts of villi in infected jird intestines and the appearance of numerous vesicles in the infected<br />

epithelium.<br />

KEY WORDS: Echinostoma trivolvis, worm expulsion, worm kinetics, intestine, cytopathology, jird, Meriones<br />

unguiculatus, SEM, TEM.<br />

Echinostoma trivolvis (Cort, 1914) Kanev, contained sulfomucins, whereas those in ham-<br />

1985, is expelled within several weeks of infec- sters contained sialomucins. Probably differention<br />

from the intestines of various strains of ces in mucin characteristics account in part for<br />

mice (Mus musculus Linnaeus, 1758): ICR (Ho- differences in infectivity of E. trivolvis in these<br />

sier and Fried, 1986; Weinstein and Fried, hosts. Thus, differences in infectivity depend in<br />

1991), BALB/c (Fujino et al., 1993), C3H (Fu- part on genetic differences in murine strains<br />

jino and Fried, 1993a; Fujino et al., 1996), and used as hosts of echinostomes.<br />

Swiss Webster (Hosier and Fried, 1986) mice, Studies on E. trivolvis in jirds have not been<br />

whereas this echinostome species is retained for done. However, Christensen et al. (1990) demore<br />

than 15 weeks in the intestines of golden scribed survival and fecundity of an allopatric<br />

hamsters (Mesocricetus auratus Waterhouse, echinostome species, Echinostoma caproni Ri-<br />

1839) (Huffman et al., 1986; Franco et al., chard, 1964, in hamsters and jirds (Meriones un-<br />

1986). Fujino et al. (1993, 1996) noted that guiculatus Milne-Edwards, 18<strong>67</strong>). Mahler et al.<br />

worm expulsion is mainly caused by an in- (1995) noted considerable differences in the recreased<br />

secretion of mucins by hyperplastic gob- productive capacity of E. caproni in hamsters<br />

let cells. Fujino and Fried (1993b; 1996) exam- versus jirds. The hamster was more susceptible<br />

ined glycoconjugates in intestinal mucins in to E. caproni infection than was the jird.<br />

C3H mice versus golden hamsters infected with The purpose of the present study was to report<br />

E. trivolvis and reported that goblet cells in mice information on the infection, growth, and distribution<br />

of E. trivolvis in jirds and to compare our<br />

data with previous studies on this species in<br />

5 Corresponding author. mouse strains and in the golden hamster. Path-<br />

236<br />

Copyright © 2011, The Helminthological Society of Washington


Table 1. Infectivity and distribution of Echinostoma trivolvis in jirds.<br />

Group<br />

A<br />

B<br />

C<br />

D<br />

E<br />

F<br />

Day<br />

postinfection<br />

5<br />

8<br />

10<br />

12<br />

15<br />

17<br />

No. of<br />

exposed<br />

(infected)<br />

jirds<br />

7 (7)<br />

7 (7)<br />

7 (7)<br />

7 (5)<br />

7 (1)<br />

7 (0)<br />

* I = anterior; II = middle; III = posterior.<br />

t Not significant.<br />

Mean (± SE)<br />

No. (%) of worms<br />

recovered<br />

14.1 ± 2.4 (35.4)<br />

13.6 ± 2.2 (33.9)t<br />

4.3 ± 0.9 (10.7)<br />

1.7 ± 1.8 (4.3)<br />

0. 1 ± 0.4 (0.4)<br />

0<br />

ological, histochemical, and electron microscopical<br />

studies of the intestines of jirds infected<br />

with E. trivolvis were also carried out.<br />

Materials and Methods<br />

Metacercarial cysts of Echinostoma trivolvis were<br />

obtained from the kidney and pericardial sac of laboratory-infected<br />

Biomphalaria glabrata (Say, 1816)<br />

snails. The worm strain was previously described by<br />

Fujino and Fried (1993a). Forty cysts were fed via a<br />

stomach tube to each jird, and 7 jirds were lightly<br />

anesthetized with ether and killed by cervical dislocation<br />

at days 5, 8, 10, 12, 15, and 17 postinfection<br />

(p.i.). Six groups of untreated control jirds, 7 per<br />

group, were also killed on the same days as the infected<br />

hosts. The jirds were starved for about 12 hr<br />

prior to necropsy to avoid food residue in the intestine.<br />

The intestine was removed and opened longitudinally<br />

to determine worm location. The worms were counted,<br />

and their distribution was recorded in the small intestine,<br />

which was divided equally into anterior, middle,<br />

and posterior regions, and in the cecum and colon plus<br />

rectum. Where applicable, Student's /-test was used to<br />

analyze differences between means, and P < 0.05 was<br />

considered statistically significant.<br />

For histological samples, pieces of intestine (2 cm<br />

long) located 10 cm anterior to the cecum, corresponding<br />

to the middle jejunum to ileum, were excised and<br />

fixed for 3 hr in Carney's fixative. The samples were<br />

dehydrated with an ethanol series and embedded in<br />

paraffin. Histological sections 5ixm thick were stained<br />

with periodic-acid Schiff for goblet cell mucins. Mucosal<br />

mast cells were stained with alcian blue (pH 0.3)<br />

and safranin O. All counts were expressed as the number<br />

of cells per villus-crypt unit (VCU) (Miller and<br />

Jarrett, 1971) for goblet cells and cells per 10 VCU<br />

for mast cells. Thirty to 50 VCUs were analyzed per<br />

host. Logarithmic transformation of data: geometric<br />

means (antilog of mean log of data) was performed.<br />

For comparison of the cell counts, goblet cell numbers<br />

were multiplied 10 times as for the mast cells. This<br />

transformation tends to stabilize variance and to normalize<br />

such data.<br />

For scanning electron microscopy (SEM), the intestinal<br />

tissue from jirds infected with E. trivolvis at 8<br />

FUJINO ET AL.—EXPULSION OF ECHINOSTOMES 237<br />

No. of worms located in the:<br />

Small intestine<br />

Total (I II III)1 Cecum<br />

97 ( 4 49 44)<br />

90 (30 21 39)<br />

30(13 8 9)<br />

10 ( 3 4 3)<br />

1(1 0 0)<br />

0<br />

1<br />

4<br />

0<br />

1<br />

0<br />

0<br />

Colon +<br />

rectum<br />

and 10 days p.i. and control tissues were excised from<br />

the upper ileum, opened longitudinally with fine needles,<br />

and pinned on small rubber boards in physiological<br />

saline. The intestinal debris was removed by gentle<br />

flow of saline forced over the surface with a pipette.<br />

After a brief rinse in 0.1M sodium cacodylate buffer<br />

(pH 7.4), the specimens were fixed for 3 hr with 3%<br />

glutaraldehyde, postfixed for 3 hr in 0.1 M osmium<br />

tetroxide (pH 7.4), and then dehydrated in an ethanol<br />

series. The material was dried in a carbon dioxide critical-point<br />

drying apparatus (Hitachi HCP-2, Tokyo, Japan),<br />

coated with palladium in a Hitachi E 1030, Tokyo,<br />

Japan ion sputter, and examined in a Hitachi s-<br />

450 SEM, Tokyo, Japan at 10 kV. For transmission<br />

electron microscopy (TEM), the material was prepared<br />

as described for SEM procedures in Fujino and Fried<br />

(1993a). Ultrathin sections stained with uranyl acetate<br />

and lead acetate were viewed in a JEOL JEM 1210<br />

electron microscope operating at 80 kV.<br />

Results and Discussion<br />

Infectivity and worm recovery data are presented<br />

in Table 1. All jirds were infected with<br />

E. trivolvis at days 5, 8, and 10 p.i., and this was<br />

confirmed by fecal examination under the microscope.<br />

By day 12 p.i., 5 of 7 jirds were infected,<br />

but only 1 was infected by day 15 p.i.<br />

Worm recoveries were 35.4% and 33.9% at days<br />

5 and 8 p.i., respectively, and this difference was<br />

not statistically significant. The recovery data<br />

dropped to 10.7% at day 10 p.i., fell markedly<br />

to 4.3% by day 12 p.i., and finally to 0.4% by<br />

day 15 p.i. Most worms were expelled between<br />

days 10 and 15 p.i., and all were expelled by<br />

day 17 p.i. Most worms were found in the middle<br />

to posterior part of the small intestine at day<br />

5 p.i. The worms moved mainly anteriad to the<br />

middle of the small intestine by day 10 p.i. Such<br />

an anteriad worm shift was reported previously<br />

in ICR mice infected with E. trivolvis at day 21<br />

p.i. (Weinstein and Fried, 1991). In the present<br />

Copyright © 2011, The Helminthological Society of Washington<br />

1<br />

0<br />

0<br />

1<br />

0<br />

0


238 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

500<br />

0.01<br />

8 10<br />

Day post-infection<br />

Figure 1. Goblet cell (mean ± SE)/VCU (•) and mast cell (•) (mean ± SE)/VCU numbers of the<br />

anterior section of the ileum of jirds, each of which was infected with Echinostoma trivolvis metacercarial<br />

cysts. A logarithmic transformation was performed on the number of the cells for normalizing the data.<br />

For comparison of the cell counts, goblet cell numbers were multiplied 10 times as for the mast cells.<br />

study, the worms moved posteriad to the cecum<br />

and colon plus rectum by day 12 p.i. In BALE/<br />

c mice infected with E. trivolvis, the recovery<br />

rate of the worms was over 44% for days 6-10<br />

p.i. and worm expulsion occurred from day 10<br />

to 12 p.i., corresponding to the peak increase in<br />

goblet cells (Fujino et al., 1996). Those worm<br />

recovery rates were much higher than what is<br />

seen in the present study on jirds, i.e., 35.4%<br />

and 33.9% at days 5 and 8 p.i., respectively.<br />

Therefore, worm expulsion occurred from days<br />

8 to 12 p.i. These data showed that worm expulsion<br />

in jirds occurred earlier than in murine<br />

hosts, probably reflecting a difference in the<br />

peak number of goblet cells in jirds and mice.<br />

Christensen et al. (1990) examined the establishment,<br />

survival, and fecundity in E. caproni and<br />

the allopatric species of E. trivolvis in hamsters<br />

and jirds. They noted that the jird exhibited an<br />

overall low susceptibility to E. caproni infection.<br />

The jird's low susceptibility to E. caproni<br />

is different from that of E. trivolvis. According<br />

to Ellerman and Morrison-Scott (1951), the jird<br />

Copyright © 2011, The Helminthological Society of Washington<br />

(M. unguiculatus) belongs to the subfamily Gerbillinae<br />

of the family Muridae and differs both<br />

taxonomically and genetically from the golden<br />

hamster (M. auratus) of the subfamily Cricetinae<br />

and also from various mouse strains of Mus<br />

musculus of the subfamily Murinae. It is known<br />

that Gerbillinae is genetically closer to Murinae<br />

than Cricetinae (Ellerman and Morrison-Scott,<br />

1951). The present infection data on E. trivolvis<br />

in jirds generally correspond to the above-noted<br />

taxonomic and genetic differences in murine<br />

hosts. In conclusion, the recoveries of E. trivolvis<br />

from jirds were lower than those from mice<br />

and much lower than those from golden hamsters.<br />

It is possible that these differences in recoveries<br />

reflect the genetic differences among<br />

these 3 hosts, jirds, mice, and hamsters.<br />

Kinetic changes in the number of goblet cells/<br />

VCU at the anterior sections (n = 50) of the<br />

ileum with or without the parasites present are<br />

shown in Figure 1. The number of goblet cells<br />

in infected jirds increased markedly, peaked at<br />

24.3 ± 0.6/VCU at day 10 p.i. and then de-<br />

12


FUJINO ET AL.—EXPULSION OF ECHINOSTOMES 239<br />

Figure 2. SEM of the control and infected intestinal surface of jirds. a. Control (normal) intestine,<br />

with round villi having regularly arranged microvilli. Scale bar = 5.0 (Jim. b. Intestine infected for 10<br />

days with Echinostoma trivolvis. The intestinal microvilli appear irregularly arranged (arrows) and partly<br />

peeled off (arrowheads). Scale bar = 5.0 fjim.<br />

clined. The number of goblet cells in the control<br />

was 9.6 ± 0.4/VCU. The numbers of mucosal<br />

mast cells were so small that their kinetic changes<br />

were examined in 10 VCU, although the<br />

number of mast cells in infected jirds increased<br />

gradually from 0.2 ± 0.42/10 VCU at day 5 p.i.<br />

to reach a peak of 1.1 ± 0.32/10 VCU at day<br />

10 p.i. and then declined gradually. For comparison<br />

of the cell counts, goblet cell numbers<br />

were multiplied 10 times as for the mast cells.<br />

The numbers of cells were log-transformed to<br />

normalize the data in Figure 1. Fujino et al.<br />

(1993) examined the worm kinetics and intestinal<br />

cytopathology in conventional and congenitally<br />

athymic BALB/c mice and noted that worm<br />

rejection was caused by goblet cell hyperplasia<br />

and not by mast cells.<br />

The SEM observations of the surface of the<br />

intestinal villi infected with E. trivolvis and the<br />

control showed a rough and irregular arrangement<br />

of microvilli in the infected intestine compared<br />

with the regular microvilli arrangement in<br />

the control intestine (Fig. 2). The intestine infected<br />

with worms was partly damaged by day<br />

10 p.i., and its epithelial surface was eroded. The<br />

TEM observations showed that the intestinal epithelium<br />

appeared more electron-dense than that<br />

in the control (not shown). The distal ends of<br />

the villi were partly broken and the microvilli<br />

were partially eroded. Numerous vesicles of various<br />

sizes appeared in the intestinal epithelium.<br />

The appearance of these vesicles was also re-<br />

ported in BALB/c and C3H mice infected with<br />

E. trivolvis by Fujino et al. (1993) and Fujino<br />

and Fried (1993a), respectively. Matrices of<br />

many small rounded mitochondria were granular<br />

and irregularly condensed. Elongate nuclei with<br />

an irregular peripheral margin had heterochromatin<br />

arranged in small patches and randomly<br />

distributed. Fujino and Fried (1996) noted histopathological<br />

differences in mouse versus hamster<br />

small intestine infected with E. trivolvis and<br />

showed no marked histopathological and histochemical<br />

changes in the hamster intestines. They<br />

suggested that the response of the hamster to E.<br />

trivolvis infection was relatively weak and that<br />

this host showed only a limited capacity to expel<br />

E. trivolvis.<br />

Literature Cited<br />

Christensen, N. 0., P. E. Simonsen, A. B. Odaibo,<br />

and H. Mahler. 1990. Establishment, survival<br />

and fecundity in Echinostoma caproni (Trematoda)<br />

infections in hamsters and jirds. Journal of the<br />

Hclminthological Society of Washington 57:104-<br />

107.<br />

Ellerman, J. R., and T. C. S. Morrison-Scott. 1951.<br />

Checklist of Palaearctic and Indian Mammals,<br />

1758 to 1946. British Museum (Natural History),<br />

London, England. 810 pp.<br />

Franco, J., J. E. Huffman, and B. Fried. 1986. In<br />

fectivity, growth, and development of Echinostoma<br />

revolutum (Digenea: Echinostomatidae) in the<br />

golden hamster, Mesocricetus auratus. Journal of<br />

<strong>Parasitology</strong> 72:142-147.<br />

Fujino, T., and B. Fried. 1993a. Expulsion of Echinostoma<br />

trivolvis (Cort, 1914) Kanev, 1985 and<br />

Copyright © 2011, The Helminthological Society of Washington


240 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

retention of E. caproni Richard, 1964 (Trematoda:<br />

Echinostomatidae) in C3H mice: pathological, ultrastructural,<br />

and cytochemical effects on the host<br />

intestine. <strong>Parasitology</strong> Research 79:286-292.<br />

, and . 1993b. Echinostoma caproni<br />

and E. trivolvis alter the binding of glycoconjugates<br />

in the intestinal mucosa of C3H mice as determined<br />

by lectin histochemistry. Journal of Helminthology<br />

<strong>67</strong>:179-188.<br />

, and . 1996. The expulsion of Echinostoma<br />

trivolvis from C3H mice: differences in<br />

gycoconjugates in mouse versus hamster small intestinal<br />

mucosa during infection. Journal of Helminthology<br />

70:115-121.<br />

-, H. Ichikawa, and I. Tada. 1996.<br />

Rapid expulsion of the intestinal trematodes Echinostoma<br />

trivolvis and E. caproni from C3H mice<br />

by trapping with increased goblet cell mucins. International<br />

Journal for <strong>Parasitology</strong> 26:319-324.<br />

-, and I. Tada. 1993. The expulsion of<br />

Echinostoma trivolvis: worm kinetics and intestinal<br />

cytopathology in conventional and congenitally<br />

athymic BALB/c mice. <strong>Parasitology</strong> 106:<br />

297-304.<br />

Hosier, D. W., and B. Fried. 1986. Infectivity,<br />

growth, and distribution of Echinostoma revolutum<br />

in Swiss Webster and ICR mice. Proceedings<br />

of the Helminthological Society of Washington<br />

53:173-176.<br />

Huffman, J. E., C. Michos, and B. Fried. 1986. Clinical<br />

and pathological effects of Echinostoma revolutum<br />

(Digenea: Echinostomatidae) in the golden<br />

hamster, Mesocricetus auratus. <strong>Parasitology</strong><br />

93:505-515.<br />

Mahler, H., N. 0. Christensen, and 0. Hindsbo.<br />

1995. Studies on the reproductive capacity of<br />

Echinostoma caproni (Trematoda) in hamsters and<br />

jirds. International Journal for <strong>Parasitology</strong> 25:<br />

705-710.<br />

Miller, H. R. P., and W. F. H. Jarret. 1971. Immune<br />

reactions in mucous membranes. I. Intestinal mast<br />

cell response during helminth expulsion in the rat.<br />

Immunology 20:277-288.<br />

Weinstein, M. S., and B. Fried. 1991. The expulsion<br />

of Echinostoma trivolvis and retention of Echinostoma<br />

caproni in the ICR mouse: pathological<br />

effects. International Journal for <strong>Parasitology</strong> 21:<br />

255-257.<br />

<strong>2000</strong>-2001 MEETING SCHEDULE OF THE<br />

HELMINTHOLOGICAL SOCIETY OF WASHINGTON<br />

11 October <strong>2000</strong><br />

15 November <strong>2000</strong><br />

17 January 2001<br />

14 March 2001<br />

5 May 2001<br />

George Washington University, Washington, DC (Contact<br />

Person: Ralph Eckerlin, 703-323-3234).<br />

Anniversary Dinner, Location to be announced.<br />

Nematology Laboratory, Beltsville Agricultural Research<br />

Service, USDA, Beltsville, MD (Contact Person: Lynn<br />

Carta, 301-504-8787).<br />

Naval Medical Research Center, 503 Robert Grant Avenue,<br />

Silver Spring, MD (Walter Reed Forest Glen Annex<br />

Bldg. 503) (Contact Person: Eileen Franke-Villasante,<br />

301-319-76<strong>67</strong>).<br />

Joint Meeting with the New Jersey Society for <strong>Parasitology</strong><br />

at the New Bolton Center, University of Pennsylvania,<br />

Kennett Square, PA (Contact Person: Jay Ferrell,<br />

215-898-8561).<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 241-243<br />

Effects of a High-Carbohydrate Diet on Growth of Echinostoma<br />

caproni in ICR Mice<br />

MARK R. DARAS, SUSAN SISBARRO, AND BERNARD FRIED'<br />

Department of Biology, Lafayette <strong>College</strong>, Easton, Pennsylvania 18042, U.S.A. (e-mail: friedb@lafayette.edu)<br />

ABSTRACT: The effects of a high-carbohydrate diet (HCD) on the host-parasite relationship of Echinostoma<br />

caproni Richard, 1964, in ICR mice were studied. The experimental diet was a customized HCD containing<br />

63% carbohydrates, 14% protein, 4% fat, and 19% cellulose. The control diet, a standard laboratory diet,<br />

contained 31% carbohydrate, 20% protein, 7% fat, and 42% cellulose. Thirty-six mice were each infected with<br />

35 metacercarial cysts; 18 mice were fed the HCD and the remaining mice received the control diet. Equal<br />

numbers of experimental and control mice were necropsied at 2, 3, and 4 weeks postinfection (p.i.). Comparisons<br />

of worm body area in uniformly fixed and stained worms were made at 2, 3, and 4 weeks p.i. There was no<br />

significant difference in body area in worms from each group at 2 and 3 weeks p.i. At 4 weeks p.i. the body<br />

area of worms from hosts on the HCD was significantly greater than that of worms from hosts on the control<br />

diet. The findings suggest that the HCD contributes to growth enhancement of E. caproni in ICR mice.<br />

KEY WORDS: trematodes, high-carbohydrate diet, Echinostoma caproni, ICR mice, growth.<br />

Previous studies in our laboratory have examined<br />

the effects of various experimental diets<br />

of hosts on growth and development of Echinostoma<br />

caproni Richard, 1964, in Institute for<br />

Cancer Research (ICR) mice. Sudati et al. (1996,<br />

1997) used this model to study the effects of<br />

high-lipid and high-protein diets, respectively, in<br />

ICR mice. Rosario and Fried (1999) examined<br />

the effects of a protein-free host diet on growth<br />

and development of E. caproni in ICR mice.<br />

Although studies are available on the effects<br />

of a high-carbohydrate host diet on gastrointestinal<br />

trematodes, this topic has been studied extensively<br />

in rats infected with hymenolipid cestodes<br />

(e.g., Read; 1959; Read and Simmons,<br />

1963). It is clear from the literature that hymenolipids<br />

thrive best in rodent hosts maintained on<br />

high-carbohydrate diets (see Von Brand, 1973,<br />

for review). Because of the lack of information<br />

on gastrointestinal trematodes maintained in rodent<br />

hosts fed a high-carbohydrate diet (HCD),<br />

we initiated this study to examine the effects of<br />

such a diet on worm recovery, growth, and distribution<br />

of E. caproni in ICR mice. Echinostoma<br />

caproni now is a well-established model<br />

for conducting such studies of intestinal trematode<br />

infections in nutritionally altered hosts.<br />

Gracyzk and Fried (1998) examined the recent<br />

literature on human echinostomiasis and<br />

noted that it is a common but forgotten foodborne<br />

disease. Because echinostomiasis may oc-<br />

Corresponding author.<br />

241<br />

cur in people from socioeconomic groups that<br />

have relatively high-carbohydrate, low-protein<br />

diets, studies on the effects of HCD on the model<br />

echinostome, E. caproni, seemed appropriate.<br />

Materials and Methods<br />

Metacercarial cysts of Echinostoma caproni were<br />

removed from the kidney/pericardial region of experimentally<br />

infected Biomphalaria glabrata (Say, 1818)<br />

snails and fed by stomach tube (35 cysts per mouse)<br />

to 36, 6 to 8-week-old, female ICR mice (Manger and<br />

Fried, 1993). The experimental mice were fed a customized<br />

HCD in pellet form containing 63% cornstarch<br />

as a source of carbohydrate, 14% protein, 4%<br />

fat, and 19% cellulose (Dyets Inc., Bethlehem, Pennsylvania,<br />

U.S.A.). The control mice were fed a standardized<br />

rat-mouse-hamster (RMH) 3000 diet in pellet<br />

form containing 31% carbohydrate, 20% protein, 7%<br />

fat, and 42% cellulose (US Biochemicals Co., Cleveland,<br />

Ohio, U.S.A.). Both diets contained essential vitamins<br />

and minerals as described previously. The HCD<br />

was about 1.3 times more calorific than the normal diet<br />

(Rosario and Fried, 1999).<br />

A total of 36 mice was used in the experiment; 18<br />

mice were maintained on the HCD, and the remainder<br />

on the RMH diet. On the day of infection, the mice<br />

were weighed and maintained 6 per cage on either the<br />

HCD or the RMH diet. Food and water were provided<br />

ad libitum. Six mice on the HCD and 6 mice on the<br />

RMH diet were each necropsied at 2, 3, and 4 weeks<br />

post infection (p.i.). Mice were weighed on the day<br />

they were fed cysts and at necropsy. At that time, the<br />

small intestine was removed from the pyloric sphincter<br />

to the ileocecal valve and divided into 5 equal sections<br />

numbered 1-5, beginning with the pylorus. Worms<br />

were removed from the small intestine, and their location<br />

and number in each section were recorded.<br />

Worms were rinsed in Locke's solution and fixed in<br />

hot (85°C) alcohol-formalin-acetic acid. Twenty<br />

Copyright © 2011, The Helminthological Society of Washington


242 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

0 1 2 3 4<br />

Weeks postinfection<br />

Figure 1. Mean (± SE) weights of mice on high<br />

carbohydrate (diamonds) versus control (squares)<br />

diet at 0—4 weeks postinfection.<br />

worms at each data point were selected at random from<br />

mice on the HCD and RMH diets and stained in Gower's<br />

carmine, dehydrated in ethanol, cleared in xylene,<br />

and mounted in Permount (Kaufman and Fried,<br />

1994). Length and maximum width measurements of<br />

worms were made with the aid of a calibrated ocular<br />

micrometer to give body area in mm2 for control and<br />

experimental worms at 2, 3, and 4 weeks p.i. Length<br />

and width measurements were also made on the gonads<br />

and suckers to determine if there were significant<br />

differences in organ sizes between worms on HCD<br />

versus RMH diet (Sudati et al., 1997). Whenever possible,<br />

differences in means between groups were determined<br />

using Student's f-test, with P < 0.05 being<br />

considered significant.<br />

Results<br />

Mean weights of mice on both the HCD and<br />

RMH diet are shown in Figure 1. Mouse weight<br />

in both groups increased rapidly until 2 weeks<br />

p.i. and then less rapidly until 4 weeks p.i. Although<br />

the weights of mice on the RMH diet<br />

were slightly higher than those of mice on the<br />

HCD diet, there was no significant difference in<br />

mouse weight between groups at any week p.i.<br />

There was no apparent difference in food consumption<br />

in mice on either diet.<br />

From 2 to 4 weeks p.i., the small intestines of<br />

hosts on the HCD were yellow compared to the<br />

tan-colored intestines of hosts on the RMH diet;<br />

the intestines of mice on the HCD were thinner,<br />

more translucent, and more brittle than those of<br />

hosts on the RMH diet. All worms from hosts<br />

on both diets were ovigerous at 2 to 4 weeks p.i.<br />

Copyright © 2011, The Helminthological Society of Washington<br />

2 3 4<br />

Weeks postinfection<br />

Figure 2. Effects of diet on mean (± SE) E. caproni<br />

worm body area; control diet (closed bar) and<br />

high-carbohydrate diet (open bar).<br />

Eggs taken at random from some worms maintained<br />

on the HCD, when incubated in artificial<br />

spring water, produced miracidia that were capable<br />

of infecting B. glabrata.<br />

The mean body areas of worms from the hosts<br />

on the RHM diet and on the HCD are shown in<br />

Figure 2. At 2 and 3 weeks p.i., there were no<br />

significant differences in the body areas of<br />

worms from either group. However, a significant<br />

increase in body areas was seen in worms from<br />

the experimental hosts at 4 weeks p.i. compared<br />

with that of worms from the control hosts. There<br />

was a significant difference at 4 weeks p.i. in<br />

the length of the anterior and posterior testes and<br />

in the diameter of the acetabulum and oral sucker<br />

of worms from the HCD group compared<br />

with those on the RMH diet.<br />

The percent worm recovery is shown in Figure<br />

3 and was similar in control and experimental<br />

mice at all 3 sampling points with about 50%<br />

recovery in both groups at all data points. More<br />

worms from hosts on the RMH diet were located<br />

in segments 3 and 4 than those from hosts on<br />

the HCD at sampling points. Considerably more<br />

worms on the HCD were located in segment 5,<br />

compared with worms on the RMH diet at all<br />

sampling points. Worms from the HCD group<br />

were more widely dispersed in their hosts than<br />

those from hosts on the RMH diet; worms from<br />

HCD hosts were also located more posteriad<br />

than those from hosts on the RMH diet.


0)<br />

0><br />

O 20-<br />

O<br />

2 3 4<br />

Weeks post infect ion<br />

Figure 3. Effects of diet on E. caproni worm recovery<br />

in mice exposed to 35 cysts/hosts; control<br />

diet (closed bar) and high-carbohydrate diet (open<br />

bar).<br />

Discussion<br />

Worms from hosts on the HCD, when compared<br />

with those from mice on the RMH diet,<br />

showed a marked increase in body area at 4<br />

weeks p.i. This is the first report that documents<br />

enhanced growth of a digenean maintained in an<br />

experimental vertebrate host fed an HCD. Echinostomes<br />

on the HCD showed greater body area<br />

by 4 weeks p.i. Reasons for the increase in<br />

worm body area are not readily apparent from<br />

the findings in this study. We have no way of<br />

knowing if the HCD had a direct effect on worm<br />

growth (i.e., if the worms consumed more carbohydrates<br />

from the HCD than the RMH diet)<br />

or had an indirect effect by altering gut constituents.<br />

Our results suggest that the intestines of<br />

hosts on the HCD showed a loss of normal integrity.<br />

The HCD at 4 weeks p.i. could have<br />

contributed to the altered gut that allowed for a<br />

large number of mucosal epithelial cells to be<br />

sloughed off, thereby increasing the food supply<br />

available to the echinostomes in hosts on the<br />

HCD. Perhaps such an increased food supply<br />

DARAS ET AL.—GROWTH OF ECH1NOSTOMA CAPRONI 243<br />

was a factor in the enhanced worm growth.<br />

Since there were no control uninfected mice on<br />

the HCD, there is also no way of knowing if<br />

some of the changes in host guts may not have<br />

been caused by interactions between diet and<br />

worms.<br />

Distribution data are interesting in that, in<br />

hosts on the HCD, worms were more spread out<br />

and also located more posteriad than worms<br />

from hosts on the RMH diet. The disparate arrangement<br />

of the worms in the HCD hosts is<br />

similar to previous observations on mice infected<br />

with E. caproni and maintained on diets with<br />

altered amounts of fats and proteins (Sudati et<br />

al., 1996, 1997; Rosario and Fried, 1999).<br />

Literature Cited<br />

Graczyk, T. K., and B. Fried. 1998. Echinostomiasis:<br />

a common yet forgotten food borne disease.<br />

American Journal of Tropical Medicine and Hygiene<br />

58:501-504.<br />

Kaufman, A. R., and B. Fried. 1994. Infectivity,<br />

growth, distribution and fecundity of a six versus<br />

twenty-live metacercarial cyst inoculum of Echinostoma<br />

caproni in ICR mice. Journal of Helminthology<br />

68:203-206.<br />

Manger, P. M., Jr., and B. Fried. 1993. Infectivity,<br />

growth and distribution of preovigerous adults of<br />

Echinostoma caproni in ICR mice. Journal of Helminthology<br />

<strong>67</strong>:158-160.<br />

Read, C. P. 1959. The role of carbohydrates in the<br />

biology of cestodes. VIII. Some conclusions and<br />

hypotheses. Experimental <strong>Parasitology</strong> 8:365-<br />

382.<br />

, and J. E. Simmons. 1963. Biochemistry and<br />

physiology of tapeworms. Physiological Reviews<br />

43:263-305.<br />

Rosario, C., and B. Fried. 1999. Effects of a proteinfree<br />

diet on worm recovery, growth and distribution<br />

of Echinostoma caproni in ICR mice. Journal<br />

of Helminthology 73:1<strong>67</strong>-169.<br />

Sudati, J. E., A. Reddy, and B. Fried. 1996. Effects<br />

of high fat diets on worm recovery, growth and<br />

distribution of Echinostoma caproni in ICR mice.<br />

Journal of Helminthology 70:351-354.<br />

, F. Rivas, and B. Fried. 1997. Effects of a<br />

high protein diet on worm recovery, growth and<br />

distribution of Echinostoma caproni in ICR mice.<br />

Journal of Helminthology 71:351-354.<br />

Von Brand, T. 1973. Biochemistry of Parasites, 2nd<br />

ed. Academic Press, New York. 499 pp.<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 244-249<br />

Surface Ultrastructure of Larval Gnathostoma cf. binucleatum from<br />

Mexico<br />

MASATAKA KoGA,1-6 HIROSHIGE AKAHANE,2 RAFAEL LAMOTHE-ARGUMEDO,3<br />

DAVID OSORIO-SARABIA,3 LUIS GARC1A-PRIETO,3 JUAN MANUEL MARTINEZ-CRUZ,4<br />

SYLVIA PAz DiAZ-CAMACHO,5 AND KANAMI NOD A1<br />

1 Department of Microbiology (<strong>Parasitology</strong>), Graduate School of Medical Sciences, Kyushu University,<br />

Fukuoka 812-8582, Japan (e-mail: masakoga@linne.med.kyushu-u.ac.jp),<br />

2 Department of <strong>Parasitology</strong>, School of Medicine, Fukuoka University, Fukuoka 814-0180, Japan,<br />

3 Laboratorio de Helmintologia, Departamento de Zoologia, Institute de Biologia, Universidad Nacional<br />

Autonoma de Mexico 04510 D.F., Mexico,<br />

4 Pedro Garcia No. 918, Tierra Blanca, Veracruz, Mexico, and<br />

5 Facultad de Ciencias Quirnico-Biologicas, Universidad Autonoma de Sinaloa, Culiacan, Sinaloa, Mexico<br />

ABSTRACT: We examined the morphology of gnathostome larvae obtained in Temazcal and Sinaloa, Mexico,<br />

mainly using scanning electron microscopy. The mean body length was 4.<strong>67</strong> mm. The head had 4 transverse<br />

rows of hooklets, and the mean number of each row was 40, 44, 47, and 50. The bodies were wholly covered<br />

with minute cuticular spines along their transverse striations. The mean number of striations varied from 227 to<br />

275. The cervical papillae were situated between the 13th and 17th transverse striations, and most specimens<br />

had them between the 14th and 15th transverse striations. An excretory pore was also located between the 24th<br />

and 28th transverse striations. We identified this Mexican gnathostome as Gnathostoma cf. binucleatum Almeyda-Artigas,<br />

1991.<br />

KEY WORDS: Gnathostoma cf. binucleatum, scanning electron microscopy, morphology, Mexico.<br />

Gnathostomiasis is an important parasitic zoonosis,<br />

mainly endemic in such countries as Japan,<br />

Thailand, and Vietnam, where people often<br />

eat raw freshwater fish. For this reason, this<br />

food-borne disease was thought to be limited to<br />

Southeast Asian countries. In 1970, however, a<br />

case of human gnathostomiasis was reported in<br />

Mexico (Pelaez and Perez-Reyes, 1970). The patient<br />

was neither a traveler nor an immigrant<br />

from Southeast Asia. After this initial discovery,<br />

the number of gnathostomiasis patients increased<br />

drastically; more than 1,000 cases have<br />

been diagnosed in Mexico. The endemic area in<br />

Mexico includes 6 states, which are roughly divided<br />

into 3 regions, including the Pacific coast<br />

(Culiacan), Atlantic coast areas (Tampico), and<br />

regions (Veracruz) adjacent to Central American<br />

countries (Ogata et al., 1998). Lamothe-Argumedo<br />

et al. (1989) and Almeyda-Artigas (1991)<br />

examined the morphology of gnathostome larvae<br />

from fish in Oaxaca-Veracruz. Later Akahane<br />

et al. (1994) examined by light microscopy<br />

the morphology of the larvae collected from pelicans<br />

in the same area.<br />

We herein report the morphology of specimens<br />

of Gnathostoma cf. binucleatum Almeyda-<br />

6 Corresponding author.<br />

244<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Artigas, 1991, from Mexico, which were examined<br />

using scanning electron microscopy<br />

(SEM). The results were compared with our previous<br />

SEM study of larvae of Gnathostoma spinigerum<br />

Owen, 1836, Gnathostoma doloresi<br />

Tubangui, 1925, and Gnathostoma hispidum<br />

Fedtschenko, 1872, obtained in Japan, China,<br />

and Thailand (Koga et al., 1987, 1988, 1994).<br />

Materials and Methods<br />

Three American white pelicans (Pelecanus erythrorhynchos<br />

Gmelin, 1789) were collected in the Presidente<br />

Miguel Aleman Reservoir in Temazcal, Oaxaca,<br />

Mexico, and their muscles were examined for gnathostome<br />

larvae. The muscles were removed, chopped<br />

into small pieces, and then cut into thin slices. The<br />

slices were then placed between 2 glass plates (10 X<br />

10 cm, 2 mm thick), pressed by hand, and examined<br />

under a dissecting microscope. The muscle remnants<br />

were then digested in artificial gastric juice (0.2 g pepsin<br />

in 0.7 ml HC1/100 ml distilled water) for 3 hours<br />

at 37°C to collect any larvae that might have been<br />

overlooked. The muscles of another ichthyophagous<br />

bird, a great egret (Egrctta alba Linnaeus, 1758), captured<br />

at a dike of the San Lorenzo River in Culiacan,<br />

were also examined. These larvae were processed for<br />

morphological examination by both light microscopy<br />

and SEM. Paraffin sections of specimens were prepared<br />

by conventional methods and stained with Mayer's<br />

hematoxylin and eosin.<br />

For the SEM specimen preparations, 10 viable lar-


vae from Temazcal and 3 from Culiacan were washed<br />

in distilled water and stored in a refrigerator until the<br />

worms relaxed completely. They were then fixed in<br />

10% formalin for 7 days. The larvae then were washed<br />

overnight in running tap water to remove the fixative<br />

and were transferred to distilled water. The specimens<br />

were rinsed twice in Millonig's phosphate buffer and<br />

postfixed overnight in 0.5% OsO4 in the same buffer.<br />

All specimens were then carefully and gradually dehydrated<br />

in an ascending series of ethanol, since such<br />

specimens often shrink or have surface wrinkles because<br />

of rapid dehydration. They were transferred into<br />

amyl acetate and CO2 critical-point dried with a Hitachi<br />

HCP-2 dryer (Tokyo, Japan). The specimens<br />

were sputter-coated with gold and examined with a<br />

JEOL JSM-U3 SEM (Tokyo, Japan) operated at 15 kV.<br />

Results<br />

As many as 570 larvae were obtained from<br />

the 3 pelicans in Temazcal. Only 3 larvae were<br />

found in 5 egrets in Culiacan. The mean body<br />

length (10 larvae) was 4.<strong>67</strong> mm, measured in a<br />

relaxed state after natural death in cold distilled<br />

water. The heads had 4 transverse rows of hooklets<br />

(Fig. 1), and the mean number in each row<br />

was 40, 44, 47, and 50 booklets. The typical<br />

hooks on the head bulb had sharp tapering points<br />

composed of hard keratin that emerged from an<br />

oblong chitinous base (Fig. 2). The bodies were<br />

wholly covered with minute cuticular spines<br />

along their transverse striations. The mean number<br />

of striations varied from 227 to 275. A pair<br />

of cervical papillae was laterally situated between<br />

the 13th and 17th transverse striations<br />

(Fig. 3). In most specimens, the papillae were<br />

located between the 14th and 15th striations. A<br />

ventral excretory pore was located between the<br />

24th and 28th transverse striations (Fig. 4). A<br />

wide terminal anal opening was visible on the<br />

ventral surface, and the transverse striations on<br />

the body were limited to the extent of this opening<br />

(Fig. 5). Both ends of the larva had a pair<br />

of lateral phasmidial pores (Fig. 6).<br />

The intestinal cells had multiple nuclei in the<br />

larvae from Temazcal (Fig. 7). The larvae from<br />

Sinaloa had 2 to 7 nuclei in each intestinal cell<br />

(Fig. 8).<br />

Discussion<br />

Lamothe-Argumedo et al. (1989) determined<br />

their larval gnathostome specimens obtained<br />

from Temazcal to be Gnathostoma sp. However,<br />

based on our observations, their specimens<br />

seemed to be the same as those reported by Almeyda-Artigas<br />

(1991); both specimens of larvae<br />

were from both fish and waterfowl in the same<br />

KOGA ET AL.—SURFACE ULTRASTRUCTURE OF GNATHOSTOMA 245<br />

endemic area of human gnathostomiasis, and the<br />

descriptions of the larval morphology were quite<br />

similar. We attributed this specimen as G. binucleatum.<br />

Lamothe-Argumedo et al. (1989)<br />

had previously observed larvae in Oaxaca, Temazcal,<br />

Mexico. We think that their SEM observations<br />

were insufficient, especially regarding<br />

the location of excretory pores and numbers of<br />

the transverse striations on the larval bodies. We<br />

reexamined the Temazcal specimens using SEM<br />

and made some new observations. We also examined<br />

the surface structures of the specimens<br />

from Sinaloa, Culiacan. Previously, 5 specimens<br />

from Sinaloa were examined by Camacho et al.<br />

(1998) using SEM. They mentioned the numbers<br />

of booklets of 4 rows on the head bulb as 39,<br />

42, 44, and 49. Furthermore, they recognized 1<br />

pair of cervical papillae located between the<br />

13th and 15th striations of the cuticular spines<br />

on a single larva. The number of transverse striations<br />

on the body was more than 200. There<br />

were no descriptions regarding the location of<br />

the excretory pore. The locations of the cervical<br />

papillae, the excretory pore, and the number of<br />

transverse striations are very important for the<br />

identification of species of gnathostome larvae.<br />

As shown in Table 1, the number of transverse<br />

striations is more than 200 in G. spinigerum.<br />

However, the number is less than 200 in most<br />

specimens of G. doloresi. On the other hand, the<br />

cervical papillae and excretory pores of G. hispidum<br />

were situated more anteriorly than those<br />

of the other 2 species.<br />

In the present study, we compared the larvae<br />

from 2 districts in Mexico, Temazcal and Culiacan,<br />

and found no differences between them in<br />

the larval morphology (Table 1). In particular,<br />

the surface ultramorphologies were very similar.<br />

However, when our findings were compared<br />

with those of G. spinigerum in Thailand (Table<br />

1), they were the same, including the shape of<br />

the larval hooks, which had oblong chitinous bases<br />

and are known to be one of the characteristic<br />

structures of G. spinigerum (Miyazaki, 1960).<br />

Akahane et al. (1994) also compared the number<br />

of booklets in each row on the head bulb of the<br />

Temazcal larvae and the larvae of G. spinigerum<br />

in Thailand by light microscopy and concluded<br />

that the numbers of booklets in Temazcal larvae<br />

were slightly less than those of G. spinigerum.<br />

The intestinal epithelium of Temazcal specimens<br />

consisted of a single layer of intestinal<br />

cells, and each columnar cell had 2 to 5 nuclei<br />

Copyright © 2011, The Helminthological Society of Washington


246 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Figures 1-4. Scanning electron micrographs of Gnathostoma cf. binucleatum. 1. Lateral view of the<br />

head bulb of Temazcal specimen. The arrow indicates the cervical papilla. Scale = 50 |xm. 2. An enlarged<br />

view of the booklets. The base of each booklet has an oblong shape. Sharp keratin hooks armed posteriorly.<br />

Scale = 10 u.m. 3. A mammary form of the cervical papilla (CP) protruding from the tegument. Scale =<br />

3 jxm. 4. The oval-shaped opening of the excretory pore (EP), which opens ventrally. Scale = 3 u.m.<br />

(Akahane et al., 1994). This feature closely resembles<br />

that of the Sinaloan specimen. Once<br />

again, no differences were observed in intestinal<br />

cells between the larvae from Temazcal and Sinaloa,<br />

and we conclude that both should be included<br />

in the same species (G. binucleatum).<br />

Further, we could not differentiate G. binucleatum<br />

from G. spinigerum based on the number of<br />

nuclei in the intestinal cells. Most specimens of<br />

G. spinigerum from Thailand also had 2 to 4<br />

Copyright © 2011, The Helminthological Society of Washington<br />

nuclei in the intestinal cells. On the other hand,<br />

the number of nuclei in the intestinal cells of<br />

other Asian species, e.g., G. hispidurn and G.<br />

doloresi, have only 1 nucleus per cell (Akahane<br />

et al., 1994). Almeyda-Artigas' light microscopic<br />

observations of the larvae were limited regarding<br />

the number of booklets in each row and<br />

the number of nuclei in the intestinal cells. Recently<br />

Koga et al. (1999) experimentally obtained<br />

the adults of this Mexican gnathostome


KOGA ET AL.—SURFACE ULTRASTRUCTURE OF GNATHOSTOMA 247<br />

Figures 5, 6. Scanning electron micrographs of Gnathostoma cf. binucleatum. 5. The terminal end of<br />

a larva where the anal opening (AP) is clearly visible on the ventral surface of the larva, which has a<br />

crescent shape. Scale = 3.5 (Jim. 6. The terminal extremity of a larva, showing a lateral phasmidial pore<br />

(PH). Scale = 12 urn.<br />

and found that the eggs have no surface pits.<br />

Furthermore, Kuramochi et al. (unpublished<br />

data, 1999) found arrangement differences in the<br />

mitochondrial DNA of adult Thai specimens of<br />

G. spinigerum and the adult Mexican gnathostome.<br />

Although our SEM observations did not<br />

show typical differences in larval stages between<br />

these 2 species, we think that this Mexican<br />

gnathostome may be a separate species.<br />

Such designation must, however, await a more<br />

detailed analysis.<br />

Gnathostoma spinigerum was reported in Ecuador<br />

in 1981 (Ollague et al., 1981), yet their<br />

description remains unclear. The adult of this<br />

species should be re-examined more precisely.<br />

We assume that this human-infecting Latin<br />

American gnathostome may be the same as that<br />

of G. binucleatum.<br />

Figures 7, 8. Cross-sections of the larval intestines of Gnathostoma cf. binucleatum. 7. A cross section<br />

of the Temazcal larva. Multiple nuclei are evident in 1 cell. Scale = 20 |xm. 8. A cross-section of the<br />

Sinaloan larva. Arrows indicate the cells with 5 nuclei each. Scale = 20 urn.<br />

Copyright © 2011, The Helminthological Society of Washington


248 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Table 1. Morphological dimensions of the advanced third-stage larvae of species of Gnathostoma (data<br />

obtained by SEM).*<br />

No. of booklets<br />

on head bulb<br />

Gnathostoma<br />

species:<br />

Locality I II III IV<br />

G. binucleatunr.<br />

Temazcal 40 43 46 49<br />

Present specimens:<br />

Temazcal<br />

Sinaloa<br />

G. spinigerum:<br />

Thailand<br />

G. doloresi:<br />

Japan<br />

G. hispidum:<br />

China<br />

G. procyonis:<br />

U.S.A.<br />

39<br />

40<br />

40<br />

39<br />

40<br />

33<br />

44<br />

44<br />

43<br />

39<br />

41<br />

* ND = not described.<br />

37<br />

46<br />

45<br />

46<br />

36<br />

47<br />

41<br />

50<br />

49<br />

50<br />

38<br />

48<br />

45<br />

Location<br />

between transverse<br />

striations of transverse<br />

striations (No.<br />

Cervical Excretory of larvae<br />

papillae pore examined) References<br />

12th-13th about 30th 260(12) Lamothe-Argumedo et al. (1989)<br />

12th- 15th<br />

12th- 15th<br />

llth- 16th<br />

15th-19th<br />

10th- 13th<br />

ND*<br />

Acknowledgments<br />

The authors would like to thank Professor<br />

Isao Tada, Department of Microbiology (<strong>Parasitology</strong>),<br />

Graduate School of Medical Sciences,<br />

Kyushu University, for reviewing the manuscript.<br />

Thanks are due to Associate Professor<br />

Brian T. Quinn, Division of Applied Linguistics,<br />

Kyushu University, for final revision of the English.<br />

This work was supported by a Grant-in-<br />

Aid for International Scientific Research (Field<br />

Research No. 08041187) from the Ministry of<br />

Education, Science, Sports, and Culture, Japan.<br />

Literature Cited<br />

Akahane, H., R. Lamothe-Argumedo, J. M. Martinez-Cruz,<br />

D. Osorio-Sarabia, and L. Garcia-<br />

Prieto. 1994. A morphological observation of the<br />

advanced third-stage larvae of Mexican Gnathostoma.<br />

Japanese Journal of <strong>Parasitology</strong> 43:18-22.<br />

Almeyda-Artigas, R. J. 1991. Hallazgo de Gnathostoma<br />

binucleatum n. sp. (Nematoda: Spirurida) en<br />

felinos silvestres y el papel de peces dulceacuicolas<br />

y oligohalinos como vectores de la gnathostomiasis<br />

humana en la cuenca baja del rio Papaloapan,<br />

Oaxaca-Veracruz, Mexico. Anales del Institute<br />

de Ciencias del Mar y Limnologia, Universidad<br />

Nacional Autonoma de Mexico 18:137-<br />

155.<br />

Ash, L. R. 1962. Development of Gnathostoma procyonis<br />

Chandler, 1942, in the first and second intermediate<br />

hosts. Journal of Parasitolology 48:<br />

298-305.<br />

24th-28th<br />

23rd-24th<br />

22nd-28th<br />

25th-28th<br />

19th-20th<br />

ND<br />

227-225 (10)<br />

228-256 (3)<br />

225-256 (8)<br />

176-211 (10)<br />

202-216 (10)<br />

ND (15)<br />

Copyright © 2011, The Helminthological Society of Washington<br />

This report<br />

This report<br />

Koga et al. (1994)<br />

Koga and Ishii (1987)<br />

Koga et al. (1988)<br />

Ash (1962)<br />

Camacho, S. P. D., M. Zazueta-Ramos, E. Ponce-<br />

Torrecillas, I. Osuna-Ramirez, R. Castro-Velazquez,<br />

A. Elores-Gaxiola, J. Baquera-Heredia,<br />

K. Willms, H. Akahane, K. Ogata, and Y.<br />

Nawa. 1998. Clinical manifestations and immunodiagnosis<br />

of gnathostomiasis in Culiacan, Mexico.<br />

American Journal of Tropical Medicine and<br />

Hygiene 59:908-915.<br />

Koga, M., H. Akahane, Y. Ishii, and S. Kojima.<br />

1994. External morphology of the advanced thirdstage<br />

larvae of Gnathostoma spinigerum: a scanning<br />

electron microscopy. Japanese Journal of<br />

<strong>Parasitology</strong> 43:23-29.<br />

, , K. Ogata, R. Lamothe-Argumedo,<br />

D. Osorio-Sarabia, L. Garcia-Prieto, and J. M.<br />

Martinez-Cruz. 1999. Adult Gnathostoma cf.<br />

binucleatum obtained from dogs experimentally<br />

infected with larvae as an etiological agent in<br />

Mexican gnathostomiasis: external morphology.<br />

Journal of the Helminthological Society of Washington<br />

66:41-46.<br />

, J. Ishibashi, Y. Ishii, and T. Nishimura.<br />

1988. Scanning electron microscopic comparisons<br />

among the early and advanced third-stage larvae<br />

of Gnathostoma hispidum and the gnathostome<br />

larvae obtained from loaches. Japanese Journal of<br />

<strong>Parasitology</strong> 37:220-226.<br />

and Y. Ishii. 1987. Surface morphology of<br />

the advanced third-stage larvae of Gnathostoma<br />

doloresi: an electron microscopic study. Japanese<br />

Journal of <strong>Parasitology</strong> 36:231-235.<br />

Lamothe-Argumedo, R., R. L. Medina-Vences, S.<br />

Lopez-Jimenez, and L. Garcia-Prieto. 1989.<br />

Hallazgo de la forma infectiva de Gnathostoma<br />

sp., en peces de Temazcal, Oaxaca, Mexico. An-


ales del Institute de Biologfa de la Universidad<br />

Nacional Autonoma de Mexico, Series Zoologia<br />

60:311-320.<br />

Miyazaki, I. 1960. On the genus Gnathostorna and<br />

human gnathostomiasis, with special reference to<br />

Japan. Experimental <strong>Parasitology</strong> 9:338-370.<br />

Ogata, K., Y. Nawa, H. Akahane, S. P. Camacho,<br />

R. Lamothe-Argumedo, and A. Cruz-Reyes.<br />

1998. Gnathostomiasis in Mexico. American Jour-<br />

KOGA ET AL.—SURFACE ULTRASTRUCTURE OF GNATHOSTOMA 249<br />

nal of Tropical Medicine and Hygiene 58:316-<br />

318.<br />

Ollague, W., J. Ollague, A. Guevara de Veliz, S.<br />

Penaherrera, C. Von Buchwald, and J. Mancheno.<br />

1981. Gnathostomiasis humana en el Ecuador<br />

(larva migrans profunda). Nuestra Medicina<br />

6:9-23.<br />

Pelaez, D., and R. Perez-Reyes. 1970. Gnathostomiasis<br />

humana en America. Revista Latino-Americana<br />

de Microbiologia 12:83-91.<br />

Report on the Brayton H. Ransom Memorial Trust Fund<br />

The Brayton H. Ransom Memorial Trust Fund was established in 1936 to "Encourage and<br />

promote the study and advancement of the Science of <strong>Parasitology</strong> and related sciences." Income<br />

from the Trust currently provides token support of <strong>Comparative</strong> <strong>Parasitology</strong> and limited support<br />

for publication of meritorious manuscripts by authors lacking institutional or other backing. Donations<br />

or memorial contributions may be directed to the Secretary-Treasurer. Information about<br />

the Trust may be found in the following articles: Proceedings of the Helminthological Society of<br />

Washington (1936) 3:84-87; (1983) 50:200-204 and Journal of the Helminthological Society of<br />

Washington (1993) 60:144-150.<br />

Financial Report for 1999<br />

Balance on hand, January 1, 1999 $23,289.69<br />

Receipts: $1,286.37<br />

Contributions from Members of the Helminthological Society of Washington:<br />

1998 $84.00<br />

1999 (in memory of Dr. Francis G. Tromba) $50.00<br />

Interest received in 1999 $1,152.37<br />

Disbursements ($117.00)<br />

Grant to the Helminthological Society of Washington for 1999 ($50.00)<br />

Membership in the American Association for Zoological Nomenclature ($50.00)<br />

Service charges on accounts ($17.00)<br />

On hand, December 31, 1999 $24,459.69<br />

Robin N. Huettel<br />

Secretary-Treasurer<br />

USDA, CSREES, Agri-Box 2220,<br />

1400 Independence Avenue<br />

Washington, DC 20250<br />

Trustees of the Brayton H. Ransom Memorial Trust Fund<br />

J. Ralph Lichtenfels, President<br />

Robin N. Huettel, Secretary-Treasurer<br />

A. Morgan Golden<br />

Harley G. Sheffield<br />

Nancy D. Pacheco<br />

Copyright © 2011, The Helminthological Society of Washington


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 250-252<br />

Research Note<br />

Helminth Parasites in Six Species of Shorebirds (Charadrii) from<br />

Bristol Bay, Alaska, U.S.A.<br />

ALBERT G. CANARis1-3 AND JOHN M. KiNSELLA2<br />

1 University of Texas at El Paso, P.O. Box 717, Hamilton, Montana 59840, U.S.A. (e-mail:<br />

acanaris@bitterroot.net), and<br />

2 Department of Pathobiology, <strong>College</strong> of Veterinary Medicine, University of Florida, Gainesville, Florida<br />

32611, U.S.A. (e-mail: wormdwb@aol.com)<br />

ABSTRACT: Nineteen species of gastrointestinal helminth<br />

parasites were recovered from 6 species of charadriid<br />

shorebirds (Aves: Charadriiformes) from Bristol<br />

Bay, Alaska: the surfbird Aphriza virgata, the western<br />

sandpiper Calidris mauri, the rock sandpiper Calidris<br />

ptilocnemis, the whimbrel Numenius phaeopus, the<br />

northern phalarope Phalaropus lobatus, and the blackbellied<br />

plover Pluvialis squatarola. Cestode species<br />

were dominant (N = 14), followed by trematode species<br />

(N = 4) and an acanthocephalan (N = 1). No<br />

nematodes were observed. Only the cestode Aploparaksis<br />

daviesi infected more than 1 species of host, the<br />

surfbird Aphriza virgata and the northern phalarope<br />

Phalaropus lobatus. All species of helminths have<br />

been reported from birds on other continents, particularly<br />

Eurasia.<br />

KEY WORDS: Helminth parasites, Aves, Charadrii,<br />

surfbird, Aphriza virgata, western sandpiper, Calidris<br />

mauri, rock sandpiper, Calidris ptilocnemis, whimbrel,<br />

Numenius phaeopus, northern phalarope, Phalaropus<br />

lobatus, black-bellied plover, Pluvialis squatarola, littoral<br />

zone, Bristol Bay, Alaska, U.S.A.<br />

Bristol Bay, Alaska, and adjacent tundra provide<br />

significant habitat to nesting and postbreeding<br />

shorebirds (Suborder Charadrii) (Gill and<br />

Handel, 1981). The bay is shallow and shorebirds<br />

are easily observed foraging on extensive<br />

sand—mud habitats exposed at low tide. Observations<br />

and counts (A.G.C.) near the mouth of<br />

the Egegik River, Bristol Bay, over the summers<br />

of 1991, 1993, 1996, and 1997 indicated arrival<br />

and utilization of the littoral zone by large numbers<br />

of postbreeding shorebirds. There is no<br />

published information on helminth parasites of<br />

shorebirds from this important area and very little<br />

from the other localities on Bristol Bay.<br />

Schmidt and Neiland (1968) recorded 19 species<br />

of cestodes and trematodes from 5 species of<br />

shorebirds collected on Kvichak Bay, Bristol<br />

Corresponding author.<br />

250<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Bay. Deblock and Rausch (1968) reported Aploparaksis<br />

leonovi Spasskii, 1962, and Aploparakris<br />

stricta Spasskii, 1962, from the western<br />

sandpiper Calidris mauri Cabanis, 1857, from<br />

the watersheds of Bristol Bay. Van Cleave and<br />

Rausch (1950), in the only previous report of<br />

helminths from the surfbird Aphriza virgata<br />

Gmelin, 1789, recovered 2 immature specimens<br />

of the acanthocephalan Arythmorhynchus comptus<br />

Van Cleave and Rausch, 1950, near Juneau,<br />

Alaska. We examined these 2 host species and<br />

4 additional species from Bristol Bay and herein<br />

report our findings.<br />

Birds were obtained from a 6.0-km-long section<br />

of littoral zone just north of the mouth of<br />

the Egegik River, Bristol Bay, Alaska, between<br />

Bishop Creek (58°14'31"N, 157°29'43"W) and<br />

Big Creek (58°17'01"N, 157°32'25"W). All species<br />

of hosts were common except the surfbird<br />

and rock sandpiper Calidris ptilocnemis Coues,<br />

1873, which were rare. Five surfbirds A. virgata,<br />

5 western sandpipers C. mauri, 4 whimbrels Numenius<br />

phaeopus Linnaeus, 1758, and 1 C. ptilocnemis,<br />

were collected in 1996, and 5 A. virgata,<br />

10 black-bellied plovers Pluvialus squatarola<br />

Linnaeus, 1758, 10 northern phalaropes<br />

Phalaropus lobatus Linnaeus, 1758, in 1997.<br />

Birds were collected with a shotgun between<br />

July 2 and July 23 and examined within 6 hr.<br />

All internal organs were examined. The koilon<br />

of the ventriculus was removed, and both the<br />

ventriculus and proventriculus tissues were<br />

teased apart. Skin and blood were not examined<br />

for parasites.<br />

Acanthocephalans, cestodes, and trematodes<br />

were fixed and preserved in alcohol-formalinacetic<br />

acid, stained in Ehrlich's hematoxylin,<br />

cleared in methyl salicylate, and mounted in<br />

Canada balsam. Voucher specimens were deposited<br />

in the United <strong>State</strong>s National Helminthol-


CANARIS AND KINSELLA—RESEARCH NOTES 251<br />

Table 1. Helminth parasites of 6 species of shorebirds (Charadrii) from Bristol Bay, Alaska, U.S.A.<br />

Number<br />

Host and parasite infected<br />

Black-bellied plover, Pluvialis squatarola Linnaeus, 1758 (N =<br />

Anornotacnia ericetorum (Krabbe, 1869)<br />

Aploparaksis diagonalis Spasskii and Bobova, 1961<br />

Liga brevis (Linstow, 1884)<br />

Proterogynotaenia variabilix Belopol'skya, 1954<br />

Schistocephalus solidiix (Mueller, 1776)<br />

Wardium squatarolae Kornyushin, 1970<br />

Echinoparyphium reciirvatum (Linstow, 1873)<br />

Polymorphic magnus (Southwell, 1927)<br />

Aploparaksis daviesi Deblock and Rausch, 1968<br />

Echinocotyle tennis Clerc, 1906<br />

Trichocephaloides megalocephala (Krabbe, 1869)<br />

Plagiorchis morosovi Sobolev, 1946<br />

Surfbird Aphriza virgata Gmelin, 1789 (N = 10)<br />

Aploparaksis diagonalis Spasskii and Bobova, 1961<br />

Dictymetra nymphaea (Schrank, 1790)<br />

Lacunovermes sp. Ching, 1965<br />

Western sandpiper Calidris mauri Cabanis, 1857 (A' = 5)<br />

Aploparaksis leonovi Spasskii, 1961<br />

Kowalewskiella cingulifera (Krabbe, 1869)<br />

Whimbrel Numenius phaeopus Linnaeus, 1758 (A' = 4)<br />

Brachylaima fuscatum (Rudolphi, 1819)<br />

Rock sandpiper Calidris ptilocnemis Coues, 1873 (N = I)<br />

Wardium ampliitricha (Rudolphi, 1819)<br />

ogical Collection, Beltsville, Maryland, U.S.A.,<br />

accession numbers 89038-89055.<br />

Nineteen species of helminths were recovered<br />

from the 6 species of hosts. Cestode species<br />

were dominant (TV = 14), followed by trematode<br />

species (TV = 4) and an acanthocephalan (TV =<br />

1). No nematodes were observed. Each of the 6<br />

species of host was parasitized by at least 1 helminth<br />

species. Only the cestode Aploparaksis<br />

daviesi Deblock and Rausch, 1968, infected<br />

more than 1 species of host—the surfbird A. virgata<br />

and northern phalarope P. lobatus (Table<br />

1). All are new host records for Alaska. All species<br />

of helminths were previously reported from<br />

birds on other continents, particularly from Eurasia<br />

(Table 1).<br />

Generally, trematode species are dominant in<br />

marine habitats, and cestodes are dominant in<br />

freshwater environments (Bush, 1990; Canaris<br />

and Kinsella, 1998). In both our study and that<br />

10)<br />

7<br />

1<br />

5<br />

5<br />

1<br />

4<br />

6<br />

1<br />

t)\<br />

1<br />

2<br />

1<br />

3<br />

6<br />

8<br />

1<br />

2<br />

2<br />

1<br />

1<br />

Mean<br />

intensity<br />

14.9<br />

0.1<br />

1 1.7<br />

129.5<br />

0.1<br />

4.6<br />

62.0<br />

0.2<br />

0.2<br />

0.4<br />

0.1<br />

0.7<br />

10.5<br />

22.4<br />

89.3<br />

7.4<br />

2.6<br />

0.25<br />

2.0<br />

Range<br />

1-72<br />

—<br />

6-55<br />

1-1,155<br />

—<br />

1-34<br />

1-468<br />

—<br />

1-3<br />

—<br />

1-3<br />

1-56<br />

2-66<br />

—<br />

1-36<br />

1-12<br />

—<br />

—<br />

Other localities<br />

Europe<br />

Russia<br />

Eurasia<br />

Russia<br />

Eurasia, North America<br />

Eurasia<br />

Cosmopolitan<br />

Russia<br />

Alaska, U.S.A.<br />

Russia<br />

Eurasia<br />

Russia<br />

Africa, Eurasia<br />

Eurasia<br />

British Columbia,<br />

Canada<br />

Russia<br />

Eurasia, Guadeloupe<br />

Australia, Europe,<br />

North America,<br />

Russia<br />

Europe, North<br />

America, Russia<br />

by Schmidt and Neiland (1968), cestode species<br />

were dominant (72% and 79%, respectively).<br />

This may reflect the hosts' recent association<br />

with the terrestrial (freshwater) nesting area, an<br />

absence of proper intermediate molluscan hosts<br />

for trematodes in Bristol Bay, or both. Also, it<br />

may reflect early summer season examination of<br />

hosts in both studies. In this study, the bulk of<br />

the trematodes was obtained later in July. Trematodes<br />

obtained earlier in July were often immature<br />

or recently mature, as indicated by the<br />

presence of small numbers of eggs and lack of<br />

pigmentation of the eggshell. Small numbers or<br />

absence of species of acanthocephalans and<br />

nematodes in Bristol Bay have also been found<br />

in studies done in Canada on 3 species of shorebirds:<br />

the long-billed curlew Numenius americanus<br />

Bechstein, 1812 (Goater and Bush, 1988);<br />

the American avocet Recurvirostra americana<br />

Gmelin, 1789 (Edwards and Bush, 1989); and<br />

Copyright © 2011, The Helminthological Society of Washington


252 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

the whimbrel Catoptrophorus semipalmatus<br />

Gmelin, 1789 (Bush, 1990). The absence of<br />

nematodes from the upper digestive tract in this<br />

study is also somewhat puzzling. Anderson et al.<br />

(1996) reviewed records for these nematodes in<br />

shorebirds from North and South America. As<br />

in the present study, they found no species of<br />

the genus Skrjabinoclava Sobolev, 1943, or ventricular<br />

nematodes in 15 surfbirds (A. virgatd)<br />

or 44 northern phalaropes (P. lobatus). However,<br />

species of Skrjabinoclava were common in 83<br />

black-bellied plovers (P. squatarold), 93 western<br />

sandpipers (C. mauri), and 8 whimbrels (N.<br />

phaeopus). It is possible that the intermediate<br />

hosts of these nematodes are absent in Bristol<br />

Bay or that they were not detected in our relatively<br />

small sample sizes. Part of the explanation,<br />

at least in our study, was that skin and<br />

blood were not examined for nematodes.<br />

Most helminths reported herein are not host<br />

specific (Baer, 1962; Deblock and Rausch, 1968;<br />

Schmidt and Neiland, 1968). Low overlap in<br />

helminth species may be attributed to small sample<br />

size, but it may also be influenced by specialized<br />

feeding habits of shorebirds (Storer,<br />

1971) prior to arrival from the nesting grounds<br />

and in Bristol Bay. Natural sorting out of shorebird<br />

species into preferred feeding habitats as<br />

the tide recedes, as reported by Ehrlich et al.<br />

(1988), is easily observed (A.G.C.) on the sandmud<br />

habitats of Bristol Bay.<br />

All species of shorebirds nesting at Bristol<br />

Bay migrate to more distant southern wintering<br />

localities, many to other continents. We expect<br />

that further studies will reveal more relationships<br />

of helminth species of shorebirds on Bristol<br />

Bay to those from distant localities.<br />

The littoral zone of Bristol Bay is an important<br />

postbreeding locality for many species of<br />

shorebirds. Studies of helminth communities<br />

need to be extended, in both time and location,<br />

to understand the dynamics of the helminth<br />

communities and their interactions among the<br />

many species of shorebirds during the long days<br />

but relatively short summer.<br />

We wish to thank Hilda Ching for her opinion<br />

on Lacunovermis sp. and Jerry Solie and Jerry<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Lang for their very able assistance, support, and<br />

long friendship with A.G.C.<br />

Literature Cited<br />

Anderson, R. C., P. L. Wong, and C. M. Bartlett.<br />

1996. The acuarioid and habronematoid nematodes<br />

(Acuarioidea, Habronematoidea) of the upper<br />

digestive tract of waders. A review of observations<br />

on their host and geographic distributions<br />

and transmissions in marine environments. Parasite<br />

4:303-312.<br />

Baer, J. G. 1962. Cestoda. Pages 1-63 in Zoology of<br />

Iceland. Vol. 2, Part 12. Ejnat Munksgaard, Copenhagen.<br />

Bush, A. O. 1990. Helminth communities in avian<br />

hosts: determinants of pattern. Pages 197-232 in<br />

G. W. Esch, A. O. Bush, and J. M. Aho, eds. Parasite<br />

Communities: Patterns and Processes. Chapman<br />

and Hall, New York.<br />

Canaris, A. G., and J. M. Kinsella. 1998. Helminth<br />

parasites in four species of shorebirds (Charadriidae)<br />

on King Island, Tasmania. Papers and Proceedings<br />

of the Royal Society of Tasmania 132:<br />

49-57.<br />

Deblock, S., and R. L. Rausch. 1968. Dix Aploparaksis<br />

(Cestoda) de Charadriiformes d'Alaska et<br />

quelques autres d'ailleurs. Annales de Parasitologie<br />

Humaine et Comparee 43:429-448.<br />

Edwards, D. D., and A. O. Bush. 1989. Helminth<br />

communities in avocets: importance of the compound<br />

community. Journal of <strong>Parasitology</strong> 75:<br />

225-238.<br />

Ehrlich, P. R., D. S. Dobkin, and D. Wheye. 1988.<br />

The Birder's Handbook: A Field Guide to the Natural<br />

History of North American Birds. Simon &<br />

Schuster Inc., New York. 785 pp.<br />

Gill, Jr., R. E., and C. M. Handel. 1981. Shorebirds<br />

of the eastern Bering Sea. Pages 719-738 in W.<br />

Hood and J. A. Calder, eds. The Eastern Bering<br />

Sea Shelf: Oceanography and Resources. Vol. 2D.<br />

University of Washington Press, Seattle.<br />

Goater, C. P., and A. O Bush. 1988. Intestinal helminth<br />

communities in long-billed curlews: the importance<br />

of cogeneric host-specialists. Holarctic<br />

Ecology 11:40-145.<br />

Schmidt, G. D., and K. A. Neiland. 1968. Hymenolepis<br />

(Hym.) deblocki sp. n., and records of other<br />

helminths from charadriiform birds. Canadian<br />

Journal of Zoology 46:1037-1040.<br />

Storer, R. W. 1971. Adaptive radiation of birds. Pages<br />

149-188 in D. S. Earner, J. R. King, and K. C.<br />

Parkes, eds. Avian Biology. Vol. 1. Academic<br />

Press, New York.<br />

Van Cleave, H. J., and R. Rausch. 1950. A new species<br />

of the acanthocephalan genus Aiythmorhynchus<br />

from sandpipers of Alaska. Journal of <strong>Parasitology</strong><br />

36:278-283.


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 253-254<br />

Research Note<br />

RESEARCH NOTES 253<br />

Colobomatus embiotocae (Copepoda: Philichthyidae) from Shiner<br />

Perch, Cymatogaster aggregata (Osteichthyes: Embiotocidae) in<br />

Canadian Waters<br />

SHELLEY F. JEPPS AND TIMOTHY M. GOATER'<br />

Biology Department, Malaspina University-<strong>College</strong>, Nanaimo, British Columbia, Canada V9R 5S5 (e-mail:<br />

goatert@mala.bc.ca)<br />

ABSTRACT: During an examination of the parasitic<br />

crustacean fauna of shiner perch, Cymatogaster aggregata<br />

(Embiotocidae) from eastern Vancouver Island in<br />

Nanaimo, British Columbia, Canada, the copepod, Colobomatus<br />

embiotocae Noble, Collard, and Wilkes,<br />

1969 (Philichthyidae), was noted in the sensory ducts<br />

of the preopercular cephalic canals. Prevalence and<br />

mean intensity of C. embiotocae were 59.2% and 1.36<br />

± 0.57, respectively. This parasite was also recovered<br />

from 68.4% (mean intensity = 1.62 ± 0.65) of shiner<br />

perch sampled near Bamfield Marine Station on the<br />

western coast of Vancouver Island. The high prevalence<br />

of C. embiotocae probably reflects increased<br />

transmission resulting from the aggregation behavior<br />

of the fish host. These results establish a range extension<br />

for C. embiotocae in C. aggregata to include Canadian<br />

Pacific waters.<br />

KEY WORDS: Colobomatus embiotocae, Copepoda,<br />

shiner perch, Cymatogaster aggregata, British Columbia,<br />

Canada.<br />

Members of the poecilostome family Philichthyidae<br />

are endoparasitic copepods that occupy<br />

the subcutaneous spaces associated with<br />

the sensory canals of the skull bones and lateral<br />

line of marine fishes (Kabata, 1979). They are<br />

highly specialized parasitic copepods, with pronounced<br />

sexual dimorphism and females exhibiting<br />

reduced organs of attachment, reduced appendages,<br />

and bizarre morphological processes<br />

projecting from their bodies.<br />

The richest genus of this family, Colobomatus,<br />

is recorded from a diversity of marine teleosts<br />

and elasmobranchs (Kabata, 1979; West,<br />

1992). Colobomatus embiotocae Noble, Collard,<br />

and Wilkes, 1969, was first described from shiner<br />

perch, Cymatogaster aggregata Gibbons,<br />

1854, and was found infecting several other spe-<br />

Corresponding author.<br />

cies of embiotocid fishes in California and<br />

Oregon in the United <strong>State</strong>s and in Mexico (Noble<br />

et al., 1969). Samples were not collected<br />

from Canadian waters, though the range of C.<br />

aggregata, among the most widely distributed<br />

embiotocid fish species, extends from Port<br />

Wrangel, Alaska, U.S.A., to Quintin Bay, Baja<br />

California, Mexico (Odenweller, 1975). Arai et<br />

al. (1988) did not find C. embiotocae during<br />

their study of metazoan parasites of C. aggregata<br />

from British Columbia. To date, the only<br />

species of Colobomatus recorded from Canadian<br />

waters is Colobomatus kyphosus Sekerak, 1970,<br />

from Sebastodes alutus Gilbert, 1890, and several<br />

species of Sebastes (Sekerak, 1970; Sekerak<br />

and Arai, 1977; Kabata, 1988).<br />

Females and males of C, embiotocae have 11<br />

body segments; in the female the fourth and fifth<br />

are fused. The average length for females and<br />

males is approximately 3.7 mm and 1.2 mm, respectively<br />

(Noble et al., 1969). Diagnostic morphological<br />

features distinguishing the female<br />

parasite from other species of Colobomatus include<br />

the caudal furcae with a spine on their<br />

inside lateral surfaces, the egg-laying apparatus<br />

with a bulbous structure equipped with a flagellate<br />

seta, and 3 eyes arranged in a compact cluster.<br />

Males of C. embiotocae are distinguished on<br />

the basis of their 6-segmented first antennae and<br />

1-segmented mandibles (Noble et al., 1969).<br />

During an investigation of the parasitic crustacean<br />

fauna of C. aggregata from Piper's Lagoon,<br />

Nanaimo, British Columbia, males and females<br />

of C. embiotocae were noticed infecting<br />

the sensory canals of the skull. A total of 76 C.<br />

aggregata was seined from the littoral region<br />

during March 1996, returned to the laboratory,<br />

and killed in concentrated anesthetic (MS-222),<br />

and their cephalic sensory canals and lateral<br />

Copyright © 2011, The Helminthological Society of Washington


254 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

lines were carefully examined. Live males and<br />

females of C. embiotocae were teased out of the<br />

canals with fine needles. The prevalence and<br />

mean ± SD intensity were 59.2% and 1.36 ±<br />

0.57, respectively. There was no association between<br />

host size and copepod intensity (n = 45,<br />

r = 0.09, P = 0.557). Females were less prevalent<br />

(22.4% females vs. 48.7% males) and<br />

abundant (mean intensity of 1.0 ± 0 females vs.<br />

1.19 ± 0.46 males) than males. There was no<br />

significant difference in the intensity of males<br />

compared with the intensity of females in infected<br />

hosts (F,,54 = 2.99, P = 0.09).<br />

Colobomatus embiotocae were also present in<br />

C. aggregata caught in a trawl (March 1996) in<br />

Trevor Channel on the western coast of Vancouver<br />

Island near the Bamfield Marine Station,<br />

British Columbia. Nineteen fish were necropsied<br />

for the presence of C. embiotocae in the cephalic<br />

canals. The prevalence was 68.4% and the mean<br />

intensity was 1.62 ± 0.65. The aggregating behavior<br />

characteristic of this fish species may be<br />

one factor explaining the high prevalence of this<br />

parasite, because host aggregation likely increases<br />

contact with C. embiotocae larvae.<br />

Only 1 of the 95 fish examined from both localities<br />

had 2 female C. embiotocae sharing the<br />

same canal. The presence of a gravid female<br />

within the cephalic sensory canals may prevent<br />

or inhibit other females from establishing themselves<br />

within such a space-constrained microhabitat.<br />

The 2 females were found aligned head<br />

to furca in the left preopercular canal. All females<br />

were recovered from either the left or<br />

right preopercular canals. Males were found in<br />

all of the skull's sensory canals. Only 1 male<br />

was recovered from the lateral line, and several<br />

males were observed exiting the fish via the<br />

pores associated with the sensory canals.<br />

These observations establish the first definitive<br />

record of C. embiotocae in Canadian eastern<br />

Pacific waters and add another species to the diverse<br />

list of shiner perch parasites in Canada<br />

(Margolis and Arthur, 1979; McDonald and<br />

Margolis, 1995). Very little is known about the<br />

population biology of philichthyid copepods,<br />

probably because they are endoparasites inhabiting<br />

a unique and seldom studied microhabitat<br />

(Kabata, 1988) and are mostly found in fish species<br />

that are of limited commercial importance<br />

(West, 1992). We urge other investigators to include<br />

the sensory canals and lateral line system<br />

Copyright © 2011, The Helminthological Society of Washington<br />

of marine fish as sites to routinely examine for<br />

these copepods. The site and host specificity of<br />

these unusual parasites might inspire experimental<br />

and field-based studies examining how seasonality<br />

and aspects of the fish host's behavior<br />

and ecology interact to influence the parasite's<br />

transmission dynamics.<br />

Voucher specimens have been deposited in<br />

the United <strong>State</strong>s National Parasite Collection,<br />

Bethesda, Maryland, U.S.A. (USNPC accession<br />

No. 87634). We thank Jason Lewis for assisting<br />

with fish collections and Cameron Weighill for<br />

necropsy assistance. The manuscript benefited<br />

from constructive comments by Bob Kabata and<br />

Cam Goater.<br />

Literature Cited<br />

Arai, H. P., Z. Kabata, and D. Noakes. 1988. Studies<br />

on seasonal changes and latitudinal differences in<br />

the metazoan fauna of the shiner perch, Cvmatogaster<br />

aggregata, along the west coast of North<br />

America. Canadian Journal of Zoology 66:1514-<br />

1517.<br />

Kabata, Z. 1979. Parasitic Copepoda of British Fishes.<br />

The Ray Society, London, 152:1-468.<br />

. 1988. Copepoda and Branchiura. Pages 3-123<br />

in L. Margolis and Z. Kabata, eds. Guide to the<br />

Parasites of Fishes of Canada. Part II. Crustacea.<br />

Canadian Special Publication of Fisheries and<br />

Aquatic Sciences No. 101, Department of Fisheries<br />

and Oceans, Ottawa, Canada. 184 pp.<br />

Margolis, L., and J. R. Arthur. 1979. Synopsis of<br />

the parasites of fishes of Canada. Bulletin of the<br />

Fisheries Research Board of Canada 199:1—269.<br />

McDonald, T. E., and L. Margolis. 1995. Synopsis<br />

of the Parasites of Fishes of Canada: Supplement<br />

(1978-1993). Canadian Special Publication of<br />

Fisheries and Aquatic Sciences, Department of<br />

Fisheries and Oceans, Ottawa, Canada 122:1—265.<br />

Noble, E. R., S. B. Collard, and S. N. Wilkes. 1969.<br />

A new philichthyid copepod parasitic in the mucous<br />

canals of surfperches (Embiotocidae). Journal<br />

of <strong>Parasitology</strong> 55:435-442.<br />

Odenweller, D. B. 1975. The life history of the shiner<br />

surfperch Cymatogaster aggregata Gibbons, in<br />

Anaheim Bay, California. Pages 107-115 in E. D.<br />

Lane and C. W. Hill, eds. The Marine Resources<br />

of Anaheim Bay. <strong>State</strong> of California, The Resources<br />

Agency, Department of Fish and Game,<br />

Fish Bulletin 165. 195 pp.<br />

Sekerak, A. D. 1970. Parasitic copepods of Sehastodes<br />

alutus, including Chondracanthus triventricosus<br />

sp. nov. and Colobomatus kyphosus sp. nov.<br />

Journal of the Fisheries Research Board of Canada<br />

27:1943-1960.<br />

, and H. P. Arai. 1977. Some metazoan parasites<br />

of rockfishes of the genus Sebastes from the<br />

northeastern Pacific Ocean. Syesis 10:139-144.<br />

West, G. A. 1992. Eleven new Colobomatus species<br />

(Copepoda: Philichthyidae) from marine fishes.<br />

Systematic <strong>Parasitology</strong> 23:81 — 133.


Comp. Parasitol.<br />

<strong>67</strong>(2), 2()(X) pp. 255-258<br />

Research Note<br />

Parasites of the Green Treefrog, Hyla cinerea, from Orange Lake,<br />

Alachua County, Florida, U.S.A.<br />

TARA L. CREEL,1'4 GARRY W. FOSTER,2 AND DONALD J. FORRESTER^<br />

Department of Pathobiology, <strong>College</strong> of Veterinary Medicine, University of Florida, Gainesville, Florida<br />

32611, U.S.A. (e-mails: ' tlc@ufl.edu; 2 FosterG@mail.vetmed.ufl.edu; 3 ForresterD@mail.vetmed.ufl.edu)<br />

ABSTRACT: Four species of parasites (1 trematode, 2<br />

nematodes, and 1 protozoan) were identified from 60<br />

green treefrogs, Hyla cinerea (Schneider), collected<br />

in north-central Florida, U.S.A. The most prevalent<br />

parasites were the nematode Cosmocercella haberi<br />

(Steiner) Baker and Adamson (23%) and the protozoan<br />

Opalina sp. Purkinje and Valentin (47%). The<br />

trematode, Clinostomum attemiatum Cort, had a prevalence<br />

of 2%, and the other nematode, Rhabdias sp.<br />

Stiles and Hassall, had a prevalence of 5%. Seven<br />

females and seven males were infected with C. haberi.<br />

The prevalence and intensity of C. haberi were<br />

correlated positively with host size (wet weight and<br />

snout-vent length). There was no statistically significant<br />

difference between gender and intensity of C.<br />

haberi infection. Fourteen females and 14 males were<br />

infected with Opalina sp. The prevalence of Opalina<br />

sp. was correlated negatively with host size. Both C.<br />

haberi and Opalina sp. have been reported previously<br />

from H. cinerea. The green treefrog represents a new<br />

host record for C. attemiatum and Rhabdias sp.<br />

KEY WORDS: Hyla cinerea, Hylidae, green treefrog,<br />

helminths, Trematoda, Clinostomum attemiatum, Nematoda,<br />

Cosmocercella haberi, Rhabdias sp., Protozoa,<br />

Opalina sp., prevalence, intensity, Florida, U.S.A.<br />

The green treefrog, Hyla cinerea (Schneider,<br />

1799), is a small, bright green, yellow, or greenish-gray<br />

treefrog with a sharply defined light<br />

stripe along the upper jaw and side of the body.<br />

Its North American range extends from Delaware<br />

south to the Florida Keys, west to Texas,<br />

and north to Illinois. Hyla cinerea is found predominantly<br />

near permanent water. In the southern<br />

parts of its range it breeds from March to<br />

October (Behler and King, 1997) and is probably<br />

the most abundant species of treefrog in the<br />

Gainesville region of north-central Florida (Kilby,<br />

1945).<br />

Corresponding author.<br />

255<br />

Many parasites have been reported from hylid<br />

frogs in the United <strong>State</strong>s and Canada. Walton<br />

(1946) listed primarily nematodes, trematodes,<br />

and protozoans as being parasitic in H. cinerea.<br />

Esch and Fernandez (1993) suggested factors<br />

that may influence parasite populations. Two of<br />

these included gender and host age or, as may<br />

be inferred for some animals, host size. To our<br />

knowledge there is no standard aging technique<br />

for treefrogs; however, Koller and Gaudin<br />

(1977) stated that "larger (hence older) frogs"<br />

usually have a greater species diversity and<br />

greater intensity of infections than "smaller,<br />

younger individuals." It was assumed for this<br />

study that host size is a rough indicator of host<br />

age.<br />

We are not aware of any comprehensive<br />

studies on the parasites of green treefrogs in<br />

Florida. The purpose of this study was to examine<br />

the parasites of H. cinerea in north-central<br />

Florida, to determine the prevalence and<br />

intensity of parasitic infections, and to determine<br />

whether relationships exist between gender,<br />

wet weight, snout-vent length, and parasitic<br />

infections.<br />

Sixty green treefrogs were collected from<br />

Orange Lake (29°27'20"N 082°10'20"W), about<br />

32 km southeast of Gainesville, Florida, U.S.A.<br />

Treefrogs were collected from a small stand of<br />

oak trees at the edge of the lake using the PVC<br />

pipe technique described by Boughton (1997).<br />

The PVC pipes were checked twice a week, during<br />

the day. Thirty treefrogs were collected from<br />

September to October 1998, and 30 treefrogs<br />

were collected from January to February 1999.<br />

All laboratory work was conducted at the Wildlife<br />

Disease Research Laboratory of the University<br />

of Florida's <strong>College</strong> of Veterinary Medicine.<br />

Treefrogs were killed with tricaine methane sulfonate<br />

(MS-222) following the methods of Gold-<br />

Copyright © 2011, The Helminthological Society of Washington


256 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Table 1. Prevalence, intensity, abundance, and location of parasites in 60 green treefrogs collected from<br />

Orange Lake, Alachua County, Florida, U.S.A., 1998-1999.<br />

Prevalence<br />

Intensity<br />

Parasite species Mean Range Abundance Location*<br />

Trematoda<br />

Clinostomum attenuatum^<br />

Nematoda<br />

Cosmocercella haberi<br />

Rhabdias sp.t<br />

Protozoa<br />

Opalina sp.<br />

23<br />

5<br />

47<br />

94<br />

3<br />

1-236<br />

2-5<br />

0.02<br />

21.6<br />

0.15<br />

SK<br />

CL,LI,SI,ST<br />

LU<br />

CL,LI,SI<br />

Location in host: CL = cloaca; LI = large intestine; LU = lungs; SI = small intestine; SK = skin; ST = stomach.<br />

New host record.<br />

berg et al. (1996) and dissected within 24 hours<br />

of capture. Gender, wet weight, and snout-vent<br />

length were recorded for each individual. The<br />

skin, liver, heart, lungs, esophagus, stomach,<br />

small intestine, large intestine, cloaca, bladder,<br />

and kidneys were evaluated for parasites in separate<br />

Petri dishes under a dissecting microscope.<br />

Protozoans were fixed in Zn-PVA and stained<br />

with Giemsa. The trematode was fixed in Roudabush's<br />

AFA, stained with acetocarmine, and<br />

mounted in neutral Canada balsam. Nematodes<br />

were fixed in 70% ethanol containing 10% glycerine<br />

and mounted in lactophenol for identification.<br />

Voucher specimens have been deposited<br />

in the United <strong>State</strong>s National Parasite Collection<br />

(USNPC), Beltsville, Maryland, U.S.A. The<br />

prevalence and intensity of parasites were correlated<br />

with wet weights and snout-vent lengths<br />

of H. cinerea using Pearson product moment<br />

correlations. A f-test was used to determine<br />

whether gender was related to intensity of Cosmocercella<br />

haberi (Steiner, 1924) Baker and Adamson,<br />

1977, infections (Minitab, 1998). We did<br />

not conduct statistical tests on Clinostomum attenuatum<br />

Cort, 1913, and Rhabdias sp. Stiles<br />

and Hassall, 1905, because of their low prevalences.<br />

Terminology used follows Bush et al.<br />

(1997).<br />

Thirty-one female and 29 male green treefrogs<br />

were collected from Orange Lake (mean<br />

wet weight ±SD = 3.5g±1.5g; mean snoutvent<br />

length ± SD = 4.2 cm ± 0.6 cm). The<br />

prevalences, intensities, abundances, and locations<br />

of parasites are listed in Table 1. Twentytwo<br />

treefrogs had no parasites, 22 had only<br />

Opalina sp., 8 had only C. haberi, 4 had both<br />

Copyright © 2011, The Helminthological Society of Washington<br />

C. haberi and Opalina sp., 2 had both C. haberi<br />

and Rhabdias sp., 1 had both Rhabdias sp. and<br />

Opalina sp., and 1 had both C. attenuatum and<br />

Opalina sp. No lesions were associated with the<br />

parasites.<br />

One green treefrog was infected with C. attenuatum<br />

(USNPC No. 88956) encysted under<br />

the skin on the back. We used 2 features to identify<br />

the trematode as C. attenuatum rather than<br />

C. complanatum, which also occurs in amphibians<br />

(McAllister, 1990): the body is uniform in<br />

width (rather than wider in the hindbody as in<br />

C. complanatum), and the testes and ovary are<br />

postequatorial (rather than medial as in C. complanatum)<br />

(Cort, 1913; Baer, 1933; Ukoli,<br />

1966). Yamaguti (1971) indicated that C. attenuatum<br />

is found in frogs, primarily species of the<br />

genera Bufo Laurenti, 1768, and Rana Linnaeus,<br />

1758. The definitive hosts include the great blue<br />

heron (Ardea herodias Linnaeus, 1758), American<br />

bittern (Botaurus lentiginosus Rackett,<br />

1813), green-backed heron (Butorides striatus<br />

Linnaeus, 1758), and double-crested cormorant<br />

(Phalacrocorax auritus Lesson, 1831). Hyla cinerea<br />

is a new host record for C. attenuatum.<br />

Fourteen green treefrogs were infected with<br />

C. haberi (USNPC Nos. 88959 and 88960). Cosmocercella<br />

haberi has been reported previously<br />

in H. cinerea by Steiner (1924) and Walton<br />

(1946). A voucher specimen of C. haberi from<br />

H. cinerea was collected in Arkansas and deposited<br />

in the USNPC by C. T. McAllister in<br />

1994 (USNPC No. 84259). This nematode is a<br />

fairly common parasite of hylids and has been<br />

identified in other species such as Hyla versicolor<br />

LeConte, 1825; Hyla arenicolor Cope,


1866; and Hyla wrightorum (Taylor, 1939)<br />

(Campbell, 1968; Goldberg et al., 1996). Cosrnocercella<br />

haberi was found in the stomach,<br />

small intestine, large intestine, and cloaca of H.<br />

cinerea. Seven females and 7 males were infected<br />

with the parasite. The prevalence of C.<br />

haberi was correlated positively with the wet<br />

weights (r = 0.283, P = 0.029) and snout-vent<br />

lengths (r = 0.268, P = 0.039) of H. cinerea.<br />

There was no statistically significant difference<br />

between gender and intensity of C. haberi infection.<br />

The intensity of C. haberi infection was<br />

correlated positively with the wet weights (r =<br />

0.<strong>67</strong>8, P = 0.008) and snout-vent lengths (r =<br />

0.760, P = 0.002) of the 14 infected green treefrogs.<br />

Three green treefrogs were infected with<br />

Rhabdias sp. (USNPC No. 88958). This nematode<br />

was found at low intensities in the lungs.<br />

Rhabdias spp. are considered "cosmopolitan" in<br />

reptiles and amphibians (Baker, 1978). Other hylids,<br />

such as Hyla regilla Baird and Girard,<br />

1852, and Pseudacris crucifer (Wied-Neuwied,<br />

1838) have been reported having these parasites<br />

(Koller and Gaudin, 1977; Muzzall and Peebles,<br />

1991; Yoder and Coggins, 1996). There is no<br />

record of Rhabdias sp. from H. cinerea.<br />

Twenty-eight green treefrogs were infected<br />

with Opalina sp. Purkinje and Valentin, 1835<br />

(USNPC No. 88957). Opalina obtrigonoidea orbiculata<br />

has been reported previously in the<br />

green treefrog by Walton (1946). A voucher<br />

specimen of Opalina sp. from H. cinerea was<br />

collected in Arkansas and deposited in the<br />

USNPC by C. T. McAllister in 1994 (USNPC<br />

No. 84277). Opalina sp. is a common parasite<br />

in treefrogs (McAllister, 1991). It has also been<br />

reported in Hyla avivoca Viosca, 1928; Pseudacris<br />

clarkii Baird, 1854; and Hyla chrysoscelis<br />

Cope, 1880 (McAllister, 1991; McAllister et<br />

al., 1993; Bolek and Coggins, 1998). Opalina<br />

sp. was found at high intensities in the small<br />

intestine, large intestine, and cloaca of H. cinerea.<br />

Fourteen females and 14 males were infected<br />

with this protozoan. The prevalence of<br />

Opalina sp. was correlated negatively with the<br />

wet weights (r = -0.357, P = 0.005) and snoutvent<br />

lengths (r = -0.387, P = 0.002) of H. cinerea.<br />

Schorr et al. (1990) studied the population<br />

changes of Opalina spp. and found population<br />

declines and even loss of the parasite at<br />

metamorphosis of some anurans. No change,<br />

however, was observed in others. They attribut-<br />

RESEARCH NOTES 257<br />

ed the decline or loss of Opalina spp. in some<br />

hosts to morphological and physiological changes<br />

in the host at metamorphosis. The negative<br />

correlation of host wet weights and snout-vent<br />

lengths with prevalence of Opalina sp. may<br />

therefore be due to the loss of parasites as the<br />

treefrogs increase in size or age. The intensity<br />

of Opalina sp. infection was not determined, because<br />

of the large numbers of Opalina sp. per<br />

treefrog.<br />

We thank the following for their contributions<br />

to this work: Dr. Kathryn Sieving for assistance<br />

in the development of this project; Katie L. Heggemeier<br />

and Jeremy J. Anderson for assistance<br />

in the field; Dr. Mike Kinsella for aid in identifying<br />

parasites, especially Clinostomum attenuatum,<br />

and reviewing the manuscript; Dr. Charles<br />

H. Courtney for assistance in the statistical analysis<br />

of the data; and Dr. Marilyn G. Spalding for<br />

reviewing the manuscript. This research was<br />

funded in part by the University of Florida <strong>College</strong><br />

of Agriculture and the Department of Wildlife<br />

Ecology and Conservation. This is Florida<br />

Agricultural Experiment Station Journal Series<br />

No. R-07053.<br />

Literature Cited<br />

Baer, J. G. 1933. Note sur un nouveau trematode,<br />

Clinostomum lophophallum sp. nov., avec quelques<br />

considerations generales sur la famille des<br />

Clinostomidae. Revue Suisse de Zoologie 40:317-<br />

342.<br />

Baker, M. R. 1978. Morphology and taxonomy of<br />

Rhabdias spp. (Nematoda: Rhabdiasidae) from<br />

reptiles and amphibians of southern Ontario. Canadian<br />

Journal of Zoology 56:2127-2141.<br />

, and M. L. Adamson. 1977. The genus Cosmocercella<br />

Steiner 1924 (Nematoda: Cosmocercoidea).<br />

Canadian Journal of Zoology 55:1644-<br />

1649.<br />

Behler, J. L., and F. W. King. 1997. National Audubon<br />

Society Field Guide to North American<br />

Reptiles and Amphibians. Alfred A. Knopf, New<br />

York. 743 pp.<br />

Bolek, M. G., and J. R. Coggins. 1998. Endoparasites<br />

of Cope's gray treefrog, Hyla chrysoscelis, and<br />

western chorus frog, Pseudacris t. triseriata, from<br />

southeastern Wisconsin. Journal of the Helminthological<br />

Society of Washington 65:212-218.<br />

Boughton, R. G. 1997. The use of PVC pipe refugia<br />

as a trapping technique for Hylid treefrogs. M.S.<br />

Thesis, University of Florida, Gainesville. 96 pp.<br />

Bush, A. O., K. D. Lafferty, J. M. Lotz, and A. W.<br />

Shostak. 1997. <strong>Parasitology</strong> meets ecology on its<br />

own terms: Margolis et al. revisited. Journal of<br />

<strong>Parasitology</strong> 83:575-583.<br />

Campbell, R. A. 1968. A comparative study of the<br />

parasites of certain Salientia from Pocahontas<br />

Copyright © 2011, The Helminthological Society of Washington


258 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

<strong>State</strong> Park, Virginia. Virginia Journal of Science<br />

19:13-29.<br />

Cort, W. W. 1913. Notes on the trematode genus Clinostomum.<br />

Transactions of the American Microscopical<br />

Society 32:1<strong>67</strong>-182.<br />

Esch, G. W., and J. C. Fernandez. 1993. A Functional<br />

Biology of Parasitism. Chapman and Hall,<br />

New York. 337 pp.<br />

Goldberg, S. R., C. R. Bursey, E. W. A. Gergus, B.<br />

K. Sullivan, and Q. A. Truong. 1996. Helminths<br />

from three treefrogs Hyla arenicolor, Hyla wrightorum,<br />

and Pseudacris triseriata (Hylidae) from<br />

Arizona. Journal of <strong>Parasitology</strong> 82:833-835.<br />

Kilby, J. D. 1945. A biological analysis of the food<br />

and feeding habits of two frogs, Hyla cinerea and<br />

Rana pipiens sphenocephala. Quarterly Journal of<br />

the Florida Academy of Science 8:71-104.<br />

Roller, R. L., and A. J. Gaudin. 1977. An analysis<br />

of helminth infections in Bufo boreus (Amphibia:<br />

Bufonidae) and Hyla regilla (Amphibia: Hylidae)<br />

in southern California. Southwestern Naturalist<br />

21:503-509.<br />

McAllister, C. T. 1990. Metacercaria of Clinostornum<br />

complanatum (Rudolphi, 1814) (Trematoda: Digenea)<br />

in a Texas salamander, Eurycea neotenes<br />

(Amphibia: Caudata), with comments on C. marginatum<br />

(Rudolphi, 1819). Journal of the Helminthological<br />

Society of Washington 57:69-71.<br />

. 1991. Protozoan, helminth, and arthropod parasites<br />

of the spotted chorus frog, Pseudacris clarkii<br />

(Anura: Hylidae), from north-central Texas.<br />

Journal of the Helminthological Society of Washington<br />

58:51-56.<br />

, S. E. Trauth, S. J. Upton, and D. H. Jamieson.<br />

1993. Endoparasites of the bird-voiced<br />

treefrog, Hyla avivoca (Anura: Hylidae), from Ar-<br />

Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 258-260<br />

Research Note<br />

kansas. Journal of the Helminthological Society of<br />

Washington 60:140-143.<br />

Minitab, Inc. 1998. Minitab Version 12 for Windows<br />

95. <strong>State</strong> <strong>College</strong>, Pennsylvania.<br />

Muzzall, P. M., and C. R. Peebles. 1991. Helminths<br />

of the wood frog, Rana sylvatica, and spring peeper,<br />

Pseudacris c. crucifer, from southern Michigan.<br />

Journal of the Helminthological Society of<br />

Washington 58:263-265.<br />

Schorr, M. S., R. Altig, and W. J. Diehl. 1990. Populational<br />

changes of the enteric protozoans Opalina<br />

spp. and Nyctotherus cordiformis during the<br />

ontogeny of anuran tadpoles. Journal of Protozoology<br />

37:479-481.<br />

Steiner, G. 1924. Some nemas from the alimentary<br />

tract of the Carolina treefrog (Hyla carolinensis<br />

pennant) with a discussion of some general problems<br />

of nematology. Journal of <strong>Parasitology</strong> 11:<br />

1-32.<br />

Ukoli, F. M. A. 1966. On Clinostomum tilapiae n. sp.<br />

and C. phalacrocoracis Dubois, 1931 from Ghana,<br />

and a discussion of the systematics of the genus<br />

Clinostomum Leidy, 1856. Journal of Helminthology<br />

40:187-214.<br />

Walton, A. C. 1946. Parasites of the Hylidae (Amphibia—Hylinae)<br />

V. Anatomical Record 96:592-<br />

593.<br />

Yamaguti, S. 1971. Synopsis of the Digenetic Trematodes<br />

of Vertebrates, Vol. I. Keigaku Publishing<br />

Co., Tokyo, Japan. 1074 pp.<br />

Yoder, H. R., and J. R. Coggins. 1996. Helminth<br />

communities in the northern spring peeper, Pseudacris<br />

c. crucifer Wied, and the wood frog, Rana<br />

sylvatica Le Conte, from southeastern Wisconsin.<br />

Journal of the Helminthological Society of Washington<br />

63:211-214.<br />

Atypical Specimens of Helminth Parasites (Anoplocephala perfoliata<br />

and Thelazia lacrymalis} of Horses in Kentucky, U.S.A.<br />

HEATHER D. BAIR,' EUGENE T. LYONS,U THOMAS W. SwERCZEK,2 AND SHARON C.<br />

TOLLIVER1<br />

1 Gluck Equine Research Center and 2 Livestock Disease Diagnostic Center, Department of Veterinary Science,<br />

University of Kentucky, Lexington, Kentucky 40546-0099, U.S.A. (e-mail: elyonsl@pop.uky.edu)<br />

ABSTRACT: During a survey of internal parasites in<br />

horses at necropsy at a diagnostic laboratory in Kentucky,<br />

U.S.A., in 1998, atypical specimens of 2 species<br />

were found. Two specimens of the cecal tapeworm,<br />

Corresponding author.<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Anoplocephla perfoliata, were fused at the midportion<br />

of each individual. One eyeworm, Thelazia lacrymalis,<br />

had 3 rather than the normal 2 uteri.<br />

KEY WORDS: atypical morphology, Cestoda, cecal<br />

tapeworm, Anoplocephala perfoliata, Nematoda, eyeworm,<br />

Thelazia lacrymalis, horses, Kentucky, U.S.A.


Several horses, all with unknown antiparasitic<br />

treatment, were examined at necropsy in Kentucky,<br />

U.S.A., in 1998 in a prevalence survey<br />

for various species of internal parasites. Specimens<br />

of 2 species were atypical. One was the<br />

cecal tapeworm, Anoplocephala perfoliata<br />

(Goeze, 1782) Blanchard, 1848. The other was<br />

the eyeworm, Thelazia lacrymalis (Gurlt, 1831)<br />

Raillet and Henry, 1910.<br />

The usual habitat of A. perfoliata in the horse<br />

is the large intestine, mainly the cecum. In past<br />

surveys of dead horses in Kentucky, prevalence<br />

of A. perfoliata was about 50-60% and no differences<br />

in infection with age of the horse were<br />

found (Benton and Lyons, 1994). Detrimental<br />

effects of A. perfoliata are not always evident.<br />

Some of the problems, mainly at the attachment<br />

sites of the tapeworms, are ulceration, inflammation,<br />

edema, and a resulting diphtheritic<br />

membrane (Proudman and Trees, 1999). Possible<br />

life-threatening effects attributed to A. perfoliata<br />

are intussusception, perforation, and hypertrophied<br />

small intestine (Proudman and<br />

Trees, 1999).<br />

Among 265 A. perfoliata found in a 29-yearold<br />

Thoroughbred gelding in the present study<br />

were 2 atypical specimens joined together at the<br />

midportion (Fig. 1). Possibly there had been incomplete<br />

separation of 2 eggs during embryogenesis.<br />

Alternatively, in early development,<br />

there may have been injury to 1 specimen, and<br />

the other somehow partially invaded the afflicted<br />

individual. The authors were unable to find any<br />

reference in the literature to this type of anomaly<br />

in A. perfoliata. However, several other types of<br />

abnormalities, including 1-4 extra suckers on<br />

the scolex and tri- and tetraradiate strobila, have<br />

been reported for A. perfoliata (Lyons et al.,<br />

1997).<br />

Thelazia lacrymalis uses muscid flies, e.g.,<br />

Musca autumnalis (deGeer, 1776), as intermediate<br />

hosts. Negative effects caused by 7". lacrymalis<br />

are usually limited to conjunctivitis and<br />

excessive lacrimation (Patton and McCracken,<br />

1981). In past surveys for T. lacrymalis, about<br />

40-50% of horses under 5-6 years of age were<br />

infected; older horses had much lower prevalences<br />

(Lyons et al., 1986). This eyeworm species<br />

is associated with several parts of the eyes,<br />

including the lacrimal glands, lacrimal ducts,<br />

conjunctival sac, and nictitating membrane<br />

gland plus ducts. Typically, females in most<br />

groups of nematodes have a double or bifurcate<br />

RESEARCH NOTES 259<br />

Figure 1. Anomalous Anoplocephala perfoliata:<br />

2 specimens joined in midportion, 1 smaller (S)<br />

than the other. Scale bar = 10.0 mm.<br />

reproductive system consisting of a vulva, vagina,<br />

and 2 uteri (Fig. 2A) and 2 ovaries. In the<br />

present study, 1 of 3 female T. lacrymalis recovered<br />

from the eyes of a yearling male Thoroughbred<br />

had 3 uteri (Fig. 2B). This aberration<br />

was observed by chance, because all female T.<br />

lacrymalis in the survey were examined for the<br />

presence of embryos with the aid of a compound<br />

microscope. The specimen with the 3 uteri accidentally<br />

ruptured at the location shown in the<br />

accompanying photomicrograph (Fig. 2B),<br />

which was taken to record the embryos. Later,<br />

it was realized that the presence of 3 uteri was<br />

not normal. No references could be found regarding<br />

such an anomaly in T. lacrymalis. Hyman<br />

(1951) mentioned that polydelphic female<br />

nematodes may have more than 2, and as many<br />

as 10 or 11, ovaries and uteri. This situation occurs<br />

particularly in the Physalopteridae, which<br />

are spirurids (Hyman, 1951). Thelazia spp.,<br />

while also spirurids, are in a different family.<br />

Chandler (1924) found 3 instead of the usual 2<br />

ovaries and uteri in the ascarid, Ascaris lumbricoides<br />

(Linnaeus, 1758), and considered this<br />

highly unusual.<br />

Causes of anomalies of internal parasites are<br />

Copyright © 2011, The Helminthological Society of Washington


260 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

2A<br />

Figure 2. Photomicrographs of Thelazia lacrymalis. Part of vagina (V) and uterus (U). A. Normal<br />

individual with 2 uteri. B. Abnormal individual with 3 uteri and embryos visible in each branch. Scale<br />

bar = 50 u,m.<br />

difficult to document. It is of interest that Becklund<br />

(1960) recorded an association of phenothiazine<br />

given to sheep and morphological<br />

anomalies of male Haemonchus contortus (Rudolphi,<br />

1803) Cobb, 1898.<br />

This investigation was done in connection<br />

with a project of the Kentucky Agricultural Experiment<br />

Station and is published with the approval<br />

of the director as paper No. 99-14-120.<br />

Literature Cited<br />

Becklund, W. R. 1960. Morphological anomalies in<br />

male Haemonchus contortus (Rudolphi, 1803)<br />

Cobb, 1898 (Nematoda: Trichostrongylidae) from<br />

sheep. Proceedings of the Helminthological Society<br />

of Washington 27:194-199.<br />

Benton, R. E., and E. T. Lyons. 1994. Survey in<br />

central Kentucky for prevalence of Anoplocephala<br />

perfoliata in horses at necropsy. Veterinary <strong>Parasitology</strong><br />

55:81-86.<br />

Chandler, A. C. 1924. A note on Ascaris lumbricoides<br />

Copyright © 2011, The Helminthological Society of Washington<br />

with three uteri and ovaries. Journal of <strong>Parasitology</strong><br />

10:208.<br />

Hyman, L. H. 1951. The Invertebrates, Volume 3.<br />

Acanthocephala, Aschelminthes, and Entoprocta.<br />

McGraw-Hill Book Co., Inc., New York. 572 pp.<br />

Lichtenfels, J. R. 1975. Helminths of domestic equids.<br />

Proceedings of the Helminthological Society of<br />

Washington 42 (special issue): 1-92.<br />

Lyons, E. T., S. C. Tolliver, J. H. Drudge, T. W.<br />

Swerczek, and M. W. Crowe. 1986. Eyeworms<br />

(Thelazia lacrymalis) in one to four-year-old<br />

Thoroughbreds at necropsy in Kentucky (1984-<br />

1985). American Journal of Veterinary Research<br />

47:315-316.<br />

, , K. J. McDowell, and J. H. Drudge.<br />

1997. Atypical external characteristics of Anoplocephala<br />

perfoliata in equids in central Kentucky.<br />

Journal of the Helminthological Society of<br />

Washington 64:287-291.<br />

Patton, S., and M. D. McCracken. 1981. The occurrence<br />

and effect of Thelazia in horses. Equine<br />

Practice 3:53-57.<br />

Proudman, C. J., and A. J. Trees. 1999. Tapeworms<br />

as a cause of intestinal disease in horses. <strong>Parasitology</strong><br />

Today 15:156-159.


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 261-263<br />

Anniversary Award<br />

The Helminthological Society of Washington<br />

FRANK W. DOUVRES<br />

J. Ralph Lichtenfels, right, presents the 1999 Anniversary Award to Frank W. Douvres<br />

Mr. President, Members and Guests, Ladies and Gentlemen, as Chair of the Awards Committee,<br />

I am honored to be able to present, on behalf of the Helminthological Society of Washington, the<br />

1999 Anniversary Award to an outstanding scientist, the world authority on the in vitro cultivation<br />

of nematode parasites of livestock and a friend and mentor to many in our society, Dr. Frank W.<br />

Douvres.<br />

Frank was born to immigrant parents in the Borough of Harlem in New York City, April 16,<br />

1927. He grew up there, speaking Greek at home, and received an outstanding education at Benjamin<br />

Franklin High School, where he graduated in 1943 at the age of 16, ranking third in his<br />

class, just behind his classmate Daniel Patrick Moynihan. Frank's classmates correctly predicted<br />

that Moynihan would go into politics, but they were off the mark when they predicted that Frank<br />

Douvres would become a Russian Commissar. This prediction was based on Frank's outspoken<br />

support for the Russian war effort against Germany in World War II.<br />

Frank completed 2 years of premed at Fordham University in December 1944. He transferred to<br />

the University of Maryland in January 1945, again in premed, but April 12, 1945, just before his<br />

eighteenth birthday, he enlisted in the navy as a hospital corpsman. On learning that Frank had<br />

enlisted, Germany immediately surrendered. Later, when Frank completed basic training, Japan<br />

surrendered!<br />

Frank was discharged from the Navy in 1947 and returned to the University of Maryland, where<br />

he switched his major to microbiology and received his B.S. degree in 1948, before reaching the<br />

age of 20 yr. At the University of Maryland Frank began to meet some really interesting people<br />

who called themselves parasitologists, so he decided to stay and work on a graduate degree. He<br />

was interested in ichthyology and completed his Master of Science degree at Maryland in 1951<br />

after completing a study program that included a survey of the parasites of fish.<br />

The parasitologist at Maryland was William O. Negherbon, who counted among his students<br />

Frank Tromba, T. Bonner Stewart, Conrad Yunker, Will Smith, and Les Costello. Professor Negerbon<br />

was studying rabies and he hired Frank Tromba to collect little brown bats, for which he<br />

paid $2 each. Douvres helped Tromba collect bats and along the way discovered a new stomach<br />

261<br />

Copyright © 2011, The Helminthological Society of Washington


262 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

worm, Rictularia lucifugus (Douvres, 1956) in the bats. Frank published the description of the new<br />

species in the Proceedings of the Helminthological Society of Washington. He went on to complete<br />

a Ph.D. at the University of Maryland on the microanatomy of Rictularia lucifugus under Professor<br />

Josh Brown in 1958.<br />

In 1953 Frank married Angelica "Kiki" Vlangas, whom he met in the Greek community of<br />

Baltimore. Typical of Frank, he told Kiki on their first date that he was going to marry her. After<br />

working briefly as a cook in a New York diner (and seriously considering staying in the restaurant<br />

business), Frank followed the example of his fellow graduate students Frank Tromba and Bonner<br />

Stewart and obtained a job with the U.S. Department of Agriculture. Frank was hired by Benjamin<br />

Schwartz, Chief of the Zoological Division of the Bureau of Animal Industry, who had obtained<br />

some new money for work on parasites of cattle.<br />

His first assignment was at Tifton, Georgia. Frank worked at Tifton with Harry Herlich, Bonner<br />

Stewart, and Dale Porter from 1953 until 1955 on parasites of cattle. It was at Tifton where Frank<br />

did his landmark work on "The Morphogenesis of the Parasitic Stages of Ostertagia ostertagi"<br />

the "Morphogenesis of the Parasitic Stages of Trichostrongylus axei and T. colubriformis," and<br />

"Keys to the Identification and Differentiation of the Immature Parasitic Stages of Gastrointestinal<br />

Nematodes of Cattle." These papers are standard references, still in use today.<br />

After transferring to Beltsville in 1955, Frank teamed up again with his old pal from graduate<br />

school, Frank Tromba, and with John Lucker on numerous studies on the morphogenesis and development<br />

of nematode parasites of cattle, sheep, and pigs.<br />

During the time Frank was a student at Maryland and later at Beltsville, he, like many of us,<br />

was fortunate to have available the advice and expertise of MayBelle Chitwood. Frank called her<br />

"coach."<br />

In 1959, Lou Diamond, who was then working at Beltsville, invited Frank to try some of his<br />

nematodes in Diamond's media developed for the in vitro culture of protozoa. The success that<br />

they had with these experiments changed the direction of Frank's research. For the next 25 years,<br />

Frank Douvres made breakthrough after breakthrough in successfully culturing important nematode<br />

parasites of large food animals in clear, cell-free media.<br />

In addition to Lou Diamond, Frank credits Paul Weinstein with mentoring his early in vitro<br />

cultivation work. Clearly, however, Frank Douvres became the recognized world authority on the<br />

in vitro cultivation of nematode parasites of animals. He collaborated with Frank Tromba on the<br />

cultivation and the description of developmental stages of parasites of swine, including Stephanunis<br />

dentatus and Ascaris suum, and, with John Lucker, Halsey Vegors, Don Thompson, and later Harry<br />

Herlich, Rob Rew and Lou Gasbarre, on parasites of cattle.<br />

From the late 1960s until Frank retired in 1985, he was assisted by George Malakatis, a worldclass<br />

technician with an international reputation of excellence, earned first in the navy with Bob<br />

Kuntz and Harry Hoogstraal and later at Beltsville with Frank.<br />

In the early 1980s Frank began a short but extremely productive collaboration with Joe Urban<br />

that included numerous papers, perhaps the most significant of which were (1) Douvres and Urban.<br />

1983. Factors contributing to the in vitro development of Ascaris suum from second-stage larvae<br />

to mature adults. Journal of <strong>Parasitology</strong> 69:549-558 and (2) Douvres and Urban. 1986. Development<br />

of Ascaris suum from in vivo—derived third-stage larvae to egg-laying adults in vitro.<br />

Proceedings of the Helminthological Society of Washington 53:256-262. During this period a distinguished<br />

visiting scientist from China worked with Frank and Joe and was a coauthor on several<br />

of their papers. Dr. Xu Shoutai, Chief, Shanghai Laboratory of Animal Schistosomiasis, spent a<br />

productive 6-month sabbatical with Frank learning his in vitro methods.<br />

Frank credits Dr. A. O. Foster with strong support and encouragement for the in vitro work.<br />

Others who worked with Frank and benefited from his expertise included Lou Gasbarre, Ray Fetterer,<br />

Rob Rew, and Bob Romanowski. Frank asked me to be sure to mention some of the support<br />

staff who made significant contributions to his research, including Ray Rew, Ken Goodson, and<br />

Don Thompson.<br />

Frank retired from the U.S. Department of Agriculture in December of 1985 and became an<br />

international consultant, traveling to Townsville, Australia, where he instructed Bruce Copeman's<br />

Copyright © 2011, The Helminthological Society of Washington


ANNIVERSARY AWARD 263<br />

laboratory on the in vitro cultivation of nematodes for several months prior to the 1986 International<br />

Congress of <strong>Parasitology</strong> in Brisbane.<br />

After the Australian trip, Frank settled into retirement and took up the role of grandfather, which<br />

he now plays for grandsons Christopher and Tim and daughter Nicky. Like everything else in his<br />

professional life, Frank plays the grandfather role with enthusiasm, a strong personal style, and a<br />

sense of duty and devotion.<br />

It was these same values that made Frank Douvres not just an outstanding scientist, but one of<br />

the most unforgettable personages for his colleagues and friends. At meetings, Frank could be<br />

counted on for a direct, to-the-point question meant to be provocative. Not everyone understood<br />

and appreciated this approach, but things were never dull when Frank was around.<br />

In retirement, Frank has continued to be active in his church and the National Association of<br />

Retired Federal Employees and was the local NARFE chapter president in 1995, prior to a serious<br />

illness from which a long recovery is now almost complete. Frank and Kiki have also generously<br />

supported HelmSoc through the activities of the Brayton H. Ransom Memorial Trust Fund.<br />

The Anniversary Award of the Helminthological Society of Washington is given either for scientific<br />

achievement or for service to the Society. Dr. Frank Douvres qualifies in both respects,<br />

having served in most of the offices of the society and on the editorial board. On behalf of the<br />

society, it is a great pleasure for me to present the 1999 Anniversary Award to Dr. Frank W.<br />

Douvres. Congratulations, Frank.<br />

Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 263-264<br />

666th Meeting: Beltsville Agricultural Research<br />

Center, United <strong>State</strong>s Department of Agriculture,<br />

Beltsville, Maryland, 13 October 1999.<br />

President Eric Hoberg presided over the business<br />

meeting and the scientific session, which<br />

consisted of 3 presentations: Dr. Benjamin Rosenthal<br />

provided an overview of the phylogeography<br />

of deer ticks in eastern North America,<br />

Dr. John Carroll spoke on black-legged ticks and<br />

Lyme disease in Maryland, and Dr. Eric Hoberg<br />

provided a summary of nematode parasites of<br />

ruminants in the Mackenzie Mountains. New<br />

members included Santiago Mas-Coma (Spain),<br />

Eun-Taek Han (Korea), Marie-Claude Durette-<br />

Desset (France), Pan Cangsang (People's Republic<br />

of China), Richard Botzler (U.S.A.), Austin<br />

Maclnnis (U.S.A.), and Scott Monks (Mexico).<br />

6<strong>67</strong>th Meeting: Sabang Restaurant, Wheaton,<br />

Maryland, 17 November 1999. The anniversary<br />

MINUTES<br />

Six Hundred Sixty-Sixth Through<br />

Six Hundred Seventieth Meeting<br />

J. Ralph Lichtenfels<br />

November 17, 1999<br />

dinner meeting and program were presided over<br />

by President Eric Hoberg. The slate of officers<br />

for <strong>2000</strong> was elected and installed by the membership<br />

in attendance: Dennis J. Richardson,<br />

president; Lynn K. Carta, vice president; Pat<br />

Carney, recording secretary; and Nancy Pacheco,<br />

corresponding secretary—treasurer. Willis A.<br />

Reid, Jr., and Janet W. Reid continued in office<br />

as editors. Dr. Ralph Lichtenfels introduced the<br />

recipient of the Anniversary Award, Dr. Frank<br />

Douvres. Dr. Douvres reviewed his research career,<br />

particularly his pioneering work with the in<br />

vitro culture of nematodes. Dr. Hoberg's final<br />

action as president was to turn the meeting over<br />

to the new president, Dr. Dennis Richardson. Dr.<br />

Richardson's first action was to adjourn the<br />

6<strong>67</strong>th meeting of the society and advise the<br />

membership that the next meeting would be held<br />

at the National Museum of Natural History,<br />

Smithsonian Institution, Washington, DC, on<br />

Wednesday, 19 January <strong>2000</strong>, at 1900 h, with<br />

William Moser serving as the host.<br />

Copyright © 2011, The Helminthological Society of Washington


264 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

668th Meeting: National Museum of Natural<br />

History, Smithsonian Institution, Washington,<br />

DC, 19 January <strong>2000</strong>. President Dennis Richardson<br />

presided over the business meeting,<br />

which he summarized for the membership, and<br />

reminded the membership that the 669th meeting<br />

of the society would be held at the Johns<br />

Hopkins Montgomery County Center in Maryland,<br />

with Dr. Thomas Simpson in charge of<br />

making the local arrangements. He then introduced<br />

William Moser, who chaired the scientific<br />

session, which consisted of 4 papers: the first<br />

paper, authored by Mr. Dan Holiday and Dr.<br />

Dennis Richardson and presented by Mr. Holiday,<br />

dealt with archaeoparasitology on the Chiribaya<br />

Culture of southern <strong>Peru</strong>; the second, by<br />

Dr. Jeff Bates, provided an overview of the molecular<br />

phylogeny of the Adenophorea; and the<br />

third, by Dr. Jon Norenburg, reviewed his phylogenetic<br />

studies of the phylum Nemertea. The<br />

final speaker was Dr. Duane Hope, who provided<br />

an overview of the phylogenetic relationship<br />

between the marine nematode genera Rhabdodemania<br />

and Pandolaimus. New members included<br />

Benjamin Rosenthal (U.S.A.) and Alan<br />

Fedynich (U.S.A.).<br />

669th Meeting: Johns Hopkins Montgomery<br />

County Center, 22 March <strong>2000</strong>. The business<br />

meeting was opened by the vice president, Lynn<br />

Carter, and presided over by President Dennis<br />

Richardson. President Richardson welcomed<br />

members and guests to the meeting, and a moment<br />

of silence was observed in memory of recently<br />

deceased society members James H.<br />

Turner, Bryce C. Walton, Richard M. Sayer,<br />

Francis G. Tromba, Everett L. Schiller, and Marion<br />

M. Farr. President Richardson then introduced<br />

Dr. Alan L. Scott, who chaired the scientific<br />

program, which consisted of 2 presentations.<br />

Dr. David Sullivan summarized his work<br />

on the formation and inhibition of heme poly-<br />

Copyright © 2011, The Helminthological Society of Washington<br />

mers in parasites. His presentation was followed<br />

by Dr. Christopher V. Plowe's discussion of a<br />

molecular marker for chloroquine-resistant falciparum<br />

malaria. Following the presentations<br />

and questions from the members and guests,<br />

President Richardson thanked the speakers for<br />

their informative and digestible summaries of<br />

malaria and schistosomiasis at the molecular<br />

level, and he also thanked Dr. Tom Simpson,<br />

who arranged the meeting. Finally, he reminded<br />

the membership that the last meeting of the season<br />

would be held at the New Bolton Center,<br />

University of Pennsylvania, Kennett Square, together<br />

with the New Jersey Society of Parasitologists,<br />

on 6 May <strong>2000</strong>. New members included<br />

Al Canaris (U.S.A.), Peter Hotez (U.S.A.),<br />

Nicole Havas (U.S.A.), John Janovy, Jr.<br />

(U.S.A.), and Alan Scott (U.S.A.).<br />

<strong>67</strong>0th Meeting: New Bolton Center, University<br />

of Pennsylvania, Kennett Square, with the New<br />

Jersey Society of Parasitologists, 6 May <strong>2000</strong>.<br />

The business meeting was presided over by<br />

President Richardson. Dr. Jay Farrell presided<br />

over the scientific meeting, which consisted of 3<br />

presentations. Dr. Thomas Klei discussed immunity<br />

to equine strongyle infections. His paper<br />

was followed by Dr. David Sibley's discussion<br />

of motility and invasion of Toxoplasma. The final<br />

presentation was provided by Dr. James B.<br />

Lok on the Dauer pathway in Caenorhabditis<br />

elegans as a model for regulation of infective<br />

larval development in parasitic nematodes. New<br />

members included Ian Whittington (Australia),<br />

M. Rocio Ruiz de Ybanez (Spain), Francisco Jimenez-Ruiz<br />

(U.S.A.), Glen Dappen (U.S.A.),<br />

Alan Kocan (U.S.A.), Aaron McCormick<br />

(U.S.A.), Robin LePardo (U.S.A.), Megan Collins<br />

(U.S.A.), Mike Barger (U.S.A.), Megan<br />

Ryan (U.S.A.), Tamara Cook (U.S.A.), Kashinath<br />

Ghosh (U.S.A.), and Richard Clopton<br />

(U.S.A.).


Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> p. 265<br />

Aguirre-Macedo, L., 85<br />

Akahane, H., 244<br />

Almeida, S. C. de, 210<br />

Amin, O. R., 40, 71<br />

Bailey, J., 71<br />

Bair, H. D., 218, 258<br />

Benz, G. W., 190<br />

Boczori, K., 230<br />

Bolek, M. G., 202<br />

Bolette, D. P., 114<br />

Botzler, R. G., 135<br />

Bouamer, S., 169<br />

Brooks, D. R., 1<br />

Buck, S., 135<br />

Bullard, S. A., 190<br />

Bursey, C. R., 60, 109, 118, 129<br />

Canaris, A. G., 250<br />

Champoux, L., 26<br />

Cheam, H., 118<br />

Coady, N. R., 32<br />

Coggins, J. R., 202<br />

Creel, T. L., 255<br />

Dailey, M. D., 165<br />

Daras, M. R., 241<br />

Diaz-Camacho, S. P., 244<br />

Doi, T., 224<br />

Domke, W., 71<br />

Duclos, L. M., 197<br />

Durette-Desset, M.-C1., 66<br />

Eidelman, W. S., 71<br />

Elsey, R. M., 122<br />

Emery, M. B., 133<br />

Forrester, D. J., 124, 255<br />

Foster, G. W., 124, 255<br />

Fried, B., 236, 241<br />

Fujino, T., 236<br />

Fujisaki, A., 224<br />

Comp. Parasitol.<br />

<strong>67</strong>(2), <strong>2000</strong> pp. 265-270<br />

AUTHOR INDEX FOR VOLUME <strong>67</strong><br />

Fukuda, K., 236<br />

Galicia-Guerrero, S., 129<br />

Garcia-Prieto, L., 92, 244<br />

Goater, T. M., 253<br />

Goldberg, S. R., 60, 109, 118, 129,<br />

165<br />

Hanelt, B., 107<br />

Hasegawa, H., 224<br />

Heckmann, R. A., 40<br />

Hemdal, J., 190<br />

Hoberg, E. P., 1<br />

Ichikawa, H., 236<br />

Janovy, J., Jr., 107<br />

Jepps, S. F, 253<br />

Jimenez-Ruiz, F. A., 145<br />

Joy, J. E., 133<br />

Kinsella, J. M., 124, 250<br />

Koga, M., 244<br />

Kritsky, D. C., 76, 145<br />

Ladd-Wilson, S., 135<br />

Lamothe-Argumedo, R., 244<br />

Leon-Regagnon, V., 92<br />

Lichtenfels, J. R., 189, 261<br />

Lyons, E. T., 218, 258<br />

Machado, P. M., 210<br />

Marcogliese, D. J., 26<br />

Martinez-Cruz, J. M., 244<br />

Mendoza-Franco, E., 76, 85<br />

Miyata, A., 224<br />

Moler, P. E., 124<br />

Morand, S., 169<br />

Muzzall, P. M., 181<br />

Nakano, T., 236<br />

Nguyen, V. H., 40<br />

Nickol, B. B., 32<br />

Noda, K., 244<br />

Olson, K. D., 218<br />

Osorio-Sarabia, D., 244<br />

Ouellet, M., 26<br />

Overstreet, R. M., 190<br />

Pavanelli, G. C., 210<br />

Perez-Ponce de Leon, G., 92<br />

Pfeifer, G., 71<br />

Pham, N. D., 40<br />

Pham, V. L., 40<br />

Razo-Mendivil, U., 92<br />

Reid, J. W., 189<br />

Richardson, D. J., 197<br />

Rodrigue, J., 26<br />

Rodriguez-Canul, R., 85<br />

Salgado-Maldonado, G., 129<br />

Sanchez-Alvarez, A., 92<br />

Santos, A., Ill, 66<br />

Scholz, T., 76, 85<br />

Scott, T. P., 122<br />

Sey, O., 145<br />

Shinohara, T., 236<br />

Simcik, S. R., 122<br />

Sisbarro, S., 241<br />

Smales, L. R., 51<br />

Spraker, T. R., 218<br />

Swerczek, T. W., 258<br />

Takemoto, R. M., 210<br />

Ten-ell, S. P., 124<br />

Tolliver, S. C., 218, 258<br />

Vidal-Martinez, V., 85<br />

Walser, C. M., 109<br />

Wargin, B., 230<br />

West, M., 122<br />

Williams, E. H., Jr., 190<br />

KEYWORD AND SUBJECT INDEX FOR VOLUME <strong>67</strong><br />

Abbreviata anomala, 109<br />

Abbreviata sp., 109<br />

Abomasal nematodes, 135<br />

Abundance, 26, 60, 122, 129, 181,<br />

202, 210, 255<br />

Acanthocephala, 32, 40, 60, 114,<br />

122, 124, 133, 181, 210, 250<br />

Acanthocephalorhynchoid.es cholodkowskyi,<br />

comb, n., 40<br />

Acanthocephalus dims, 181<br />

265<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Acanthopagrus berda, 145<br />

Acanthopagrus bifasciatus, 145<br />

Acanthopagrus latus, 145<br />

Acanthorhodeus fortunensis, 40<br />

Acari, 124


266 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Agamidae, 109<br />

Alaska, U.S.A., 218, 250<br />

Alligator snapping turtle, 122<br />

Amblyomma dissimile, 124<br />

Ambystoma andersoni, 92<br />

Ambystoma dumerilii, 92<br />

Ambystoma lermaensis, 92<br />

Ambystoma mexicanum, 92<br />

Ambystoma tigrinum, 92<br />

Ameloblastella gen. n., 76<br />

Ameloblastella chavarriai comb.<br />

n., 76<br />

Ameloblastella mamaevi comb, n.,<br />

76<br />

Ameloblastella platensis comb, n.,<br />

76<br />

American toad, 202<br />

American white pelican, 244<br />

Amphibia, 26, 92, 129, 133, 92,<br />

165, 202, 224, 255<br />

Anatomy, 51<br />

Ancyrocephalinae, 85<br />

Anguilliformes, 190<br />

Anomotaenia ericetorum, 250<br />

Andrias japonicus, 224<br />

Angiostoma onychodactyla sp. n.,<br />

60<br />

Angiostoma plethodontis, 133<br />

Angiostomatidae, 60<br />

Anisakiasis, 71<br />

Anoplocephala perfoliata, 258<br />

Anura, 92, 129, 165, 202, 224, 255<br />

Aphanoblastella gen. n., 76<br />

Aphanoblastella mastigatus comb.<br />

n., 76<br />

Aphanoblastella robustus comb.<br />

n., 76<br />

Aphanoblastella travassosi comb.<br />

n., 76<br />

Aphriza virgata, 250<br />

Aplectana incerta, 129<br />

Aploparaksis daviesi, 250<br />

Aploparaksis diagonalis, 250<br />

Aploparaksis leonovi, 250<br />

Aquaculture, 181, 190<br />

Argentina, 76<br />

Argyrops filamentosus, 145<br />

Argyrops spinifer, 145<br />

Arkansas, U.S.A., 122<br />

Armadillidium nasatum, 32<br />

Artnadillidium vulgare, 32<br />

Asian pond loach, 224<br />

Atheriniformes, 190<br />

Atypical morphology, 258<br />

Australia, 51, 109<br />

Australian water dragon, 109<br />

Aves, 32, 244, 250<br />

Aviary, 114<br />

Batracholandros salamandrae, 133<br />

Biodiversity, 1<br />

Biogeography, 85, 92<br />

Biomphalaria glabrata, 236, 241<br />

Biosphere, 1<br />

Black-bellied plover, 250<br />

Black-tailed prairie dog, 197<br />

Blarina brevicauda, 32<br />

Brachycoelium storeriae, 133<br />

Brachylaima fuscatum, 250<br />

Brazil, 210<br />

Brevimulticaecum tenuicolle, 122<br />

Brevitritospinus subgen. n., 40<br />

Bristol Bay, Alaska, U.S.A., 250<br />

British Columbia, Canada, 253<br />

British West Indies, 190<br />

Brook trout, 181<br />

Brown trout, 181<br />

Bufo americanus americanus, 202<br />

Bufo marinus, 92, 129<br />

Bufo marmoreus, 129<br />

Bufo valliceps, 92<br />

Bufonidae, 26, 92, 202, 129<br />

Bullfrog tadpole, 26<br />

Calidris mauri, 250<br />

Calidris ptilocnemis, 250<br />

California, U.S.A., 71, 135, 165<br />

California treefrog, 165<br />

Callorhinus ursinus, 218<br />

Calydiscoides flexuosus, 145 •<br />

Canada, 26, 253<br />

Capriniana sp., 181<br />

Capsalidae, 190<br />

Carabidae, 107<br />

Carnivora, 32<br />

Carolinensis tuffi sp. n., 66<br />

Case history, 71<br />

Catadiscus rodriguezi, 92<br />

Catfish, 76<br />

Ca the be (fish), 40<br />

Caudata, 224<br />

Cavia porcellus, 197<br />

Cecal tapeworm, 258<br />

Cenotes, 76<br />

Centrorhynchus spinosus, 124<br />

Centrorhynchus sp., 129<br />

Cephalogonimus americanus, 92<br />

Cephalouterina leoi, 60<br />

Cestoda, 109, 118, 124, 181, 197,<br />

202, 210, 250, 258<br />

Chaetodon lunula, 190<br />

Chaetodontidae, 190<br />

Channidae, 40<br />

Charadrii, 250<br />

Chilodonella sp., 181<br />

Chinchilla lanigera, 197<br />

Chlaenius prasinus, 107<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Cichla monoculus, 210<br />

Cichlasoma aureum, 85<br />

Cichlasoma callolepis, 85<br />

Cichlasoma friedrichstahli, 85<br />

Cichlasoma geddesi, 85<br />

Cichlasoma helleri, 85<br />

Cichlasoma lentiginosum, 85<br />

Cichlasoma managuense, 85<br />

Cichlasoma salvini, 85<br />

Cichlasoma sp., 85<br />

Cichlasoma trimaculatum, 85<br />

Cichlasoma urophthalmus, 85<br />

Cichlidae, 85, 210<br />

Ciliophora, 181<br />

Clinostomum attenuaturn, 255<br />

Clinostomum sp., 210<br />

Cloacinidae, 51<br />

Cobitis biwae, 224<br />

Cockroach, 32<br />

Coleoptera, 107<br />

Colobomatus embiotocae, 253<br />

Common wallaroo, 51<br />

Community structure, 202, 210<br />

Congridae, 190<br />

Connecticut, U.S.A., 197<br />

Contracaecum sp., 210<br />

Copepoda, 181, 224, 253<br />

Copper-tailed skink, 118<br />

Cosrnocercella haberi, 255<br />

Cosmocercoides variabilis, 202<br />

Cricetidae, 66<br />

Crustacea, 32, 181, 224, 253<br />

Cryptoblepharus poeciloplcurus,<br />

118<br />

Cryptobranchidae, 224<br />

Ctenophorus caudicinctus, 109<br />

Ctenophorus fordi, 109<br />

Ctenophorus isolepis, 109<br />

Ctenophorus reticulatus, 109<br />

Ctenophorus scutulatus, 109<br />

Cylindrotaenia decidua, 118<br />

Cymatogaster aggregata<br />

Cynomys ludovicianus, 197<br />

Cyprinidae, 40<br />

Cystacanth, 26, 33, 60, 114, 124,<br />

129, 133<br />

Cytopathology, 236<br />

Dactylogyridae, 76, 85<br />

Deer mouse, 32<br />

Delagoa threadfin bream, 145<br />

Demidueterospinus subgen. n., 40<br />

Diagnostic <strong>Parasitology</strong> Course, 39<br />

Dictymetra nymphaea, 250<br />

Digenea, 26, 92, 124, 165, 202,<br />

210, 236, 241, 250, 255<br />

Diplectanidae, 145<br />

Diplectanum cazauxi, 145


Diplectanum sillagonum, 145<br />

Diplectanum spp., 145<br />

Diplodus noct, 145<br />

Diplostomatidae, 26<br />

Diplostomiim (Austrodiplostomum)<br />

compactum, 210<br />

Diplostomiim sp., 26, 210<br />

Domestic mouse, 32, 197, 230, 241<br />

Domestic spiny mouse, 197<br />

Diymarchon corais couperi, 124<br />

Dwarf bearded dragon, 109<br />

Eastern indigo snake, 124<br />

Echeneidae, 190<br />

Echeneis neucratoides, 190<br />

Echinocotyle tenuis, 250<br />

Echinopaiyphium recurvatum, 250<br />

Echinostoma caproni, 241<br />

Echinostoma trivolvis, 236<br />

Echinostome metacercariae, 202<br />

Ecology, 202, 210, 218, 250<br />

Editors' Acknowledgments, 223<br />

Egretta alba, 244<br />

Eleutherodactylus rhodopis, 92<br />

Embiotocidae, 253<br />

Etnoia cyanura, 118<br />

Endohelminths, 1, 26, 32, 40, 51,<br />

60, 66, 71, 76, 85, 92, 107, 109,<br />

114, 118, 122, 124, 129, 133,<br />

135, 145, 165, 169, 181, 197,<br />

202, 210, 218, 224, 230, 236,<br />

241, 244, 250, 255, 258<br />

Epinephelus areolatus, 145<br />

Epinephelus rnorio, 190<br />

Epinephelus tauvina, 145<br />

Estado de Mexico, Mexico, 92<br />

Eiibothriurn salvelini, 181<br />

European rabbit, 197<br />

Eustrongyloides sp., 124<br />

Experimental infection, 32<br />

Extraintestinal infection, 32<br />

Eyefluke, 26<br />

Eyeworm, 258<br />

Falcaustra chelydrae, 122<br />

Falcaustra wardi, 122<br />

Ferret, 197<br />

Fibricola sp., 92<br />

Fishes, 40, 76, 85, 145, 181, 190,<br />

210, 224, 253<br />

Florida, U.S.A., 124, 190, 255<br />

French Polynesia, 118<br />

Frog, 26, 92, 165, 224, 255<br />

Gambusia xanthosoma, 190<br />

Gehyra oceanica, 118<br />

Genus revision, 40, 165, 169<br />

Gigantorhynchidae, 114<br />

Global Taxonomy Initiative, 1<br />

Glucocorticoid treatment, 230<br />

Glypthelmins californiensis, 92<br />

Glypthelmins facial, 92<br />

Glypthelmins parva, 92<br />

Glypthelmins quieta, 92<br />

Glypthelmins sp., 92<br />

Gnathostorna cf. biniicleatum, 244<br />

Gnathostoma doloresi, 224<br />

Gnathostoma sp., 124<br />

Gnathostomatidae, 124, 224<br />

Golden hamster, 197<br />

Gordioidea, 107<br />

Gordius difficilis, 107<br />

Gorgoderina attenuata, 92<br />

Gorgoderina sp., 202<br />

Grand Cayman Island, 190<br />

Great egret, 244<br />

Green treefrog, 255<br />

Growth, 241<br />

Guatemala, 85<br />

Guinea pig, 197<br />

Gulf of Mexico, 190<br />

Gyrodactylus sp., 181<br />

Haematoloechus coloradensis, 92<br />

Haematoloechus cornplexus, 92<br />

Haematoloechus illimis, 92<br />

Haematoloechus longiplexus, 92<br />

Haematoloechus medioplexus, 92<br />

Haematoloechus pulcher, 92<br />

Haematoloechus sp., 92<br />

Haemonchus contortus, 135<br />

Halipegus occidualis, 92<br />

Heligmosomoidea, 66<br />

Helminthological Society of Washington:<br />

Anniversary Award, 261<br />

Articles of Incorporation, 141<br />

Constitution and By-Laws, 138<br />

Meeting Schedule, 65, 240<br />

Membership Application, 143,<br />

Minutes of Meetings, 263<br />

Mission and Vision <strong>State</strong>ments,<br />

144, 272<br />

Helminths, 1, 26, 32, 40, 51, 60,<br />

66, 71, 76, 85, 92, 107, 109, 114,<br />

118, 122, 124, 129, 133, 135,<br />

145, 165, 169, 181, 197, 202,<br />

210, 218, 224, 230, 236, 241,<br />

244, 250, 255, 258<br />

Hemiramphidae, 145<br />

Hemiramphus marginatus, 145<br />

Hermann's tortoise, 169<br />

Heteroconger has si, 190<br />

Heteromyidae, 197<br />

High-carbohydrate diet, 241<br />

Hispid pocket mouse, 32<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Hookworms, 218<br />

Horse, 258<br />

Host specificity, 85, 190<br />

Human infection, 71<br />

Hyla arenicolor, 92<br />

Hyla cadaverina, 165<br />

Hyla cinerea, 255<br />

Hyla eximia, 92<br />

Hylidae, 92, 165, 255<br />

Hymenolepis nana, 197<br />

Hynobiidae, 60<br />

INDEX 2<strong>67</strong><br />

Ichthyobodo sp., 181<br />

Ichthyophthirus multijiliis, 181<br />

ICR mice, 241<br />

India, 145<br />

Indiana, U.S.A., 190<br />

Indo-Pacific tree gecko, 118<br />

Inducible nitric oxide, 230<br />

iNOS, 230<br />

Insecta, 114<br />

Intensity, 32, 60, 109, 118, 122,<br />

124, 129, 133, 135, 181, 197,<br />

202, 210, 250, 253, 255<br />

International Code of Zoological<br />

Nomenclature (Fourth Edition),<br />

75<br />

Intestine, 236<br />

Inventory, 1<br />

Isopoda, 32<br />

Jalisco, Mexico, 92, 129<br />

Japan, 60, 224<br />

Japanese clawed salamander, 60<br />

Japanese giant salamander, 224<br />

Japanese threadfin bream, 145<br />

Jird, 197, 236<br />

Kalicephalus appendiculatus, 124<br />

Kalicephalus inermis coronellae,<br />

124<br />

Kalicephalus rectiphilus, 124<br />

Kentucky, U.S.A., 258<br />

King soldierbream, 145<br />

Kiricephalus coarctatus, 124<br />

Kowalewskiella cingulifera, 250<br />

Kreisiella chrysocampa, 109<br />

Kreisiella lesueurii, 109<br />

Kuwait, 145<br />

Lacunovermes sp., 250<br />

iMinellodiscus furcillatus sp. n.,<br />

145<br />

Lamellodiscus spp., 145<br />

Langeronia burseyi sp. n., 165<br />

Langeronia macrocirra, 165<br />

iMngeronia parva, 165<br />

Langeronia provitellaria, 165


268 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Larva, 114, 210, 224, 230, 244<br />

Lecithodendriidae, 165<br />

Lepidodactylus lugubris, 118<br />

Leptodactylus melanonotus, 92<br />

Lepidotrema kuwaitensis sp. n.,<br />

145<br />

Lepidotrema longipenis comb, n.,<br />

145<br />

Life history, 224<br />

Liga brevis, 250<br />

Lipotyphyla, 32<br />

Littoral zone, 250<br />

Lizard, 109, 118<br />

Long-tailed chinchilla, 197<br />

Lophognathus longirostris, 109<br />

Louisiana, U.S.A., 122<br />

Loxogenes (Langeronia) macrocirra,<br />

92<br />

Lozenge-marked dragon, 109<br />

Lutjanidae, 190<br />

Lutjanus campechanus, 190<br />

Macracanthorhynchus ingens, 124<br />

Macroclemys temminckii, 122<br />

Macrocyclops albidus, 224<br />

Macropodidae, 51<br />

Macropus rohustus, 51<br />

Mallee dragon, 109<br />

Mammalia, 1, 32, 66, 135, 190,<br />

197, 218, 230, 236, 241, 258<br />

Manatee, 190<br />

Marsupalia, 51<br />

Mastigophora, 181<br />

Maxvachonia brygooi, 109<br />

Maxvachonia chabaudi, 118<br />

Mediorhynchus orientalis, 114<br />

Mediorhynchus wardi, 114<br />

Meeting Announcements, 91, 209<br />

Megalodiscus americanus, 92<br />

Meriones iinguiculatus, 197, 236<br />

Mesoccstoides sp., 202<br />

Mesocoelium brevicaecum, 60<br />

Mesocoelium rnonas, 92<br />

Mesocricetus auratus, 197<br />

Mesocyclops dissimilis, 224<br />

Metacercaria, 26<br />

Methylprednisolone, 230<br />

Metoponorthus pruinosus, 32<br />

Mexico, 76, 85, 92, 129, 244<br />

Mexico City, Mexico, 92<br />

Michigan, U.S.A., 181<br />

Michoacan, Mexico, 92<br />

Microtus ochrogaster, 32<br />

Microtus pennsylvanicus, 32<br />

Military dragon, 109<br />

Misgurnus anguillicaudatus, 224<br />

Mississippi, U.S.A., 190<br />

Mollusca, 236, 241<br />

Mongolian gerbil, 197, 236<br />

Monogenea, 85, 181, 190<br />

Monogenoidea, 76, 145<br />

Moorea, 118<br />

Morocco, 169<br />

Morphology, 40, 51, 60, 66, 71, 76,<br />

85, 92, 107, 114, 145, 165, 169,<br />

224, 244, 258<br />

Moth skink, 118<br />

Mourning gecko, 118<br />

Mouse, 32, 197, 230, 241<br />

Mudpuppy, 26<br />

Muscle, 230<br />

Mus musculus, 32, 197, 230, 241<br />

Museums for Depositing of Specimens,<br />

189<br />

Mustela nivalis, 32<br />

Mustela putorius favo, 197<br />

Myxobolus cerebralis, 181<br />

Myxozoa, 181<br />

Nebraska, U.S.A., 32, 107<br />

Necturus maculosus, 26<br />

Nematoda, 51, 60, 66, 71, 109,<br />

118, 122, 124, 129, 133, 135,<br />

169, 181, 202, 210, 218, 224,<br />

230, 255, 258<br />

Nernatodeirus odocoilei, 135<br />

Nematomorpha, 107<br />

Nemipteridae, 145<br />

Nemiptenis bipunctatus, 145<br />

Nemipterus peronii, 145<br />

Neobenedenia melleni, 190<br />

Neoechinorhynchus chrysemydis,<br />

122<br />

Neoechinorhynchus emydis, 122<br />

Neoechinorhynchus pseudemydis,<br />

122<br />

New book available, 201<br />

New geographical record, 60, 76,<br />

85, 92, 107, 109, 118, 122, 124,<br />

129, 133, 135, 145, 190, 250,<br />

253, 255<br />

New host record, 85, 92, 107, 109,<br />

118, 122, 124, 129, 133, 135,<br />

145, 190, 197, 250, 255<br />

New South Wales, Australia, 51<br />

New taxon, 40, 51, 60, 66, 76, 97,<br />

145, 165, 169<br />

Northern fur seal pups, 218<br />

Northern phalarope, 250<br />

Northern Territory, Australia, 109<br />

Norway rat, 197<br />

Notchedfin threadfm bream, 145<br />

Numenius phaeopus, 250<br />

Oaxaca, Mexico, 92, 244<br />

Obituary Notices:<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Alan F. Bird, 168<br />

Marion M. Farr, 180<br />

Michael J. Patrick, 164<br />

Everett Lyle Schiller, 31<br />

Obtuse barracuda, 145<br />

Oceanic gecko, 118<br />

Ochetosoma elongatum, 124<br />

Ochetosoma kansense, 124<br />

Ochetosoma sp., 92<br />

Ochoterenella digiticauda, 129<br />

Odocoileus hemionus fuliginatus,<br />

135<br />

Ohio, U.S.A., 190<br />

Oncorhynchus my kiss, 181<br />

Onychodactylus japonicus, 60<br />

Oochoristica piankai, 109<br />

Opalina sp., 255<br />

Ophicephalus rnaculatus, 40<br />

Oryctolagus cuniculus, 197<br />

Osteichthyes, 145, 253<br />

Oswaldocruzia pipiens, 202<br />

Otolithes argenteus, 145<br />

Oxyurid sp., 118<br />

Oxyuroids, 169<br />

Pachymedusa dachnicolor, 92<br />

Pademelon, 51<br />

Pallisentis, 40<br />

Pallisentis sensu stricto, new diagnosis,<br />

40<br />

Pallisentis tetraodontae, new synonym,<br />

40<br />

Pallisentis (Brevitritospinus) vietnamensis<br />

subgen. et sp. n., 40<br />

Pallisentis (Pallisentis) pesteri,<br />

comb, n., 40<br />

Parana River, Brazil, 210<br />

Paraphaiyngodon fitzroyi, 109<br />

Parapharyngodon japonicus, 60<br />

Pararaosentis gen. n., 40<br />

Pararaosentis golvani, comb, n.,<br />

40<br />

Pelecanus erythrorhynchos, 244<br />

Pennsylvania, U.S.A., 114<br />

Pentastoma, 124<br />

Perciformes, 210<br />

Periplaneta americana, 32<br />

Perognathus flavescens, 32<br />

Perognathus hispidus, 32<br />

Perornyscus leucopus, 32<br />

Peromyscus maniculatus, 32<br />

Perornyscus pectoralis, 66<br />

Persian Gulf, 145<br />

Petenia splendida, 85<br />

Pets, 197<br />

Phalaropus lobatus, 250<br />

Pharyngodon oceanicus, 118<br />

Pharyngodonidae, 118, 169


Philichthyidae, 253<br />

Phodopus sungorus, 197<br />

Phylogeny, 1<br />

Physaloptera obtussirna, 124<br />

Physaloptera sp., 124, 129<br />

Physocephalus sp., 129<br />

Pickhandle barracuda, 145<br />

Pimclodidae, 76<br />

Pimelodus clarias, 76<br />

Pisces, 40, 76, 85, 145, 181, 190,<br />

210, 224, 253<br />

Plagiorchis morosovi, 251<br />

Plagiorhynchus cylindraceus, 32<br />

Plethodon richmondi, 133<br />

Pluvialis squatarola, 250<br />

Poeciliidae, 190<br />

Pogona minor, 109<br />

Polymorphic magnus, 250<br />

Popovastrongylus pluteus sp. n.,<br />

51<br />

Popovastrongylus tasmaniensis sp.<br />

n., 51<br />

Popovastrongylus wallabiae, 51<br />

Prevalence, 26, 32, 60, 109, 118,<br />

122, 124, 129, 133, 135, 181,<br />

197, 202, 210, 218, 253, 255<br />

Proteocephalus macrophallus, 210<br />

Proteocephalus microscopicus, 210<br />

Proteocephalus sp., 124, 181<br />

Proterogynotaenia variabilis, 250<br />

Protolamellodiscus senilobatus sp.<br />

n., 145<br />

Protozoa, 181, 255<br />

Pseudolamellodiscus sphyraenae,<br />

145<br />

Pseudopolystoma dendriticum, 60<br />

Pseudorhabdosynochus spp., 145<br />

Pseudoterranova decipiens, 71<br />

Public aquaria, 190<br />

Puebla, Mexico, 92<br />

Puerto Rico, 190<br />

Pycnoscelis surinarnensis, 114<br />

Quadrigyridae, 40, 210<br />

Quadrigyrus machadoi, 210<br />

Quebec, Canada, 26<br />

Rainbow trout, 181<br />

Rana brownorurn, 92<br />

Rana catesbeiana, 26<br />

Rana dunni, 92<br />

Rana forreri, 92<br />

Rana megapoda, 92<br />

Rana montezumae, 92<br />

Rana neovolcanica, 92<br />

Rana nigromaculata, 224<br />

Rana rugosa, 224<br />

Rana vaillanti, 92<br />

Ransom Trust Fund Report, 249<br />

Rattus norvegicus, 197<br />

Ravine salamander, 133<br />

Red-necked wallaby, 51<br />

Red Sea seabream, 145<br />

Reithrodontomys rnegalotis, 32<br />

Reptilia, 109, 118, 122, 124, 169<br />

Rhabdias americanus, 202<br />

Rhabdias fuelleborni, 129<br />

Rhabdias sp., 255<br />

Rhamdia guatemalensis, 76<br />

Ring-tailed dragon, 109<br />

Robin, 32<br />

Rock sandpiper, 250<br />

Rodentia, 32, 66, 197, 230, 236,<br />

241<br />

St. Lawrence River, Canada, 26<br />

St. Paul Island, Alaska, U.S.A., 218<br />

Salamander, 26, 60, 92, 133, 224<br />

Salmincola edwardsii, 181<br />

Salmo trutta, 181<br />

Salmonidae, 181<br />

Salvelinus fontinalis, 181<br />

Sand loach, 224<br />

Sauria, 109, 118<br />

Scanning electron microscopy,<br />

169, 236, 244<br />

Schistocephalus solidus, 250<br />

Sciadiclethrutn bravohollisae, 85<br />

Sciadiclethrum meekii, 85<br />

Sciadiclethrum mexicanum, 85<br />

Sciadiclethrum splendidae, 85<br />

Sciadocephalus megalodiscus, 210<br />

Sciaenidac, 145<br />

Seasonal study, 202<br />

Second-stage larva, 224<br />

SEM, 169, 236, 244<br />

Serpinerna trispinosus, 122<br />

Serranidae, 145, 193<br />

Shiner perch, 253<br />

Shorebirds, 250<br />

Short-tailed shrew, 32<br />

Siberian hamster, 197<br />

Sillaginidae, 145<br />

Sillago siharna, 145<br />

Siluriformes, 76<br />

Silver sillago, 145<br />

Sinaloa, Mexico, 248<br />

Skrjabinoptera gallardi, 109<br />

Skrjabinoptera goldrnanae, 109<br />

Skrjabinoptera sp., 118<br />

Small-scaled terapon, 145<br />

Smilisca baudinii, 92<br />

Snail, 236, 241<br />

Snake, 124<br />

Snake-eyed skink, 118<br />

Snake head mullet, 40<br />

Copyright © 2011, The Helminthological Society of Washington<br />

INDEX 269<br />

Soldierbream, 145<br />

Sorex cinereus, 32<br />

Southern mule deer, 136<br />

Spain, 169<br />

Sparidae, 145<br />

Spauligodon gehyrae, 118<br />

Spennophilus tridecemlineatus, 32<br />

Sphyraena chrysotaenia, 145<br />

Sphyraena jello, 145<br />

Sphyraena obtusata, 145<br />

Sphyraenidae, 145<br />

Spiroxys hanzaki, 224<br />

Spur-thighed tortoise, 169<br />

Starling, 32<br />

Strongyloides sp., 124<br />

Stump-toed gecko, 118<br />

Sturnus vulgaris, 32<br />

Surfbird, 250<br />

Surinam cockroach, 114<br />

Survey, 1, 26, 32, 60, 76, 85, 92,<br />

109, 118, 122, 124, 129, 133,<br />

135, 145, 197, 201, 218, 250,<br />

253, 255<br />

Tadpole, 26<br />

Tasmania, Australia, 51<br />

Tasmanian pademelon, 51<br />

Taxonomic key, 40, 165, 169<br />

Taxonomy, 1, 40, 51, 60, 66, 76,<br />

85, 92, 115, 145, 165, 169<br />

Teladorsagia circumcincta, 135<br />

Teleostei, 40, 76, 85, 145, 190,<br />

210, 224, 253<br />

TEM, 236<br />

Terapon puta, 145<br />

Teraponidae, 145<br />

Terranova caballeroi, 124<br />

Testudinidae, 169<br />

Testudo graeca, 169<br />

Testudo hermanni, 169<br />

Texas, U.S.A., 66<br />

Thaparia bourgati sp. n., 169<br />

Thaparia capensis, 169<br />

Thaparia carlosfeliui sp. n., 169<br />

Thaparia contortospicula, 169<br />

Thaparia domerguei, 169<br />

Thaparia macrocephala, 169<br />

Thaparia microcephala, 169<br />

Thaparia rnacrospiculum, 169<br />

Thaparia thapari australis, 169<br />

Thaparia thapari rysavyi, 169<br />

Thaparia thapari thapari, 169<br />

Thelazia lacrymalis, 258<br />

Third-stage larva, 71, 224<br />

Thirteen-lined ground squirrel, 32<br />

Thylogale billiardierii, 51<br />

Toad, 92, 129, 202<br />

Tortoise, 169


270 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

Trachelipus rathkei, 32<br />

Transmission electron microscopy,<br />

236<br />

Trematoda, 60, 124, 165, 202, 241,<br />

250, 255<br />

Trichechus manatux, 190<br />

Trichinella pseudospiralis, 230<br />

Trichinella spi rails, 230<br />

Trichocephaloides megalocephala,<br />

250<br />

Trichodina sp., 181<br />

Trichostrongylina, 66, 135<br />

Trout, 181<br />

Truttaedacnitis sp., 181<br />

Tucunare, 210<br />

Turdus migratorius, 32<br />

Turtle, 122<br />

Ultrastructure, 169, 236, 244<br />

Uncinaria lucasi, 218<br />

Urodela, 92<br />

U.S.A., 32, 66, 71, 107, 122, 124,<br />

133, 135, 165, 181, 190, 197,<br />

202, 218, 250, 255, 258<br />

Veracruz, Mexico, 92<br />

Vietnam, 40<br />

Vivid metallic ground beetle, 107<br />

Wallaby, 51<br />

Wanaristrongyla ctenoti, 109<br />

Wardium amphitricha, 250<br />

Wardium squatarolae, 250<br />

West Virginia, U.S.A., 133<br />

Copyright © 2011, The Helminthological Society of Washington<br />

Western Australia, 109<br />

Western netted dragon, 109<br />

Western sandpiper, 250<br />

Whimbrel, 250<br />

White-ankled mouse, 66<br />

Whitefin sharksucker, 190<br />

Wisconsin, U.S.A., 202<br />

Wood mouse, 32<br />

Worm expulsion, 236<br />

Worm kinetics, 236<br />

Yellowstrip barracuda, 145<br />

Yucatan, Mexico, 76<br />

Zoogeography, 85, 92, 190<br />

Zoonosis, 197


Name:<br />

MEMBERSHIP APPLICATION 271<br />

APPLICATION FOR MEMBERSHIP<br />

in the<br />

HELMINTHOLOGICAL SOCIETY OF WASHINGTON<br />

(Please Type or Print Legibly)<br />

Mailing Address:<br />

Present Position and Name of Institution (If you are a student, write "Student" and indicate the<br />

degree you are working for and the university/department):<br />

Phone: _<br />

E-Mail:<br />

Highest Degree Earned and the Year Received:<br />

Fields of Interest:<br />

FAX:<br />

If you are established and experienced in your field, would you consent to be a reviewer for<br />

manuscripts submitted for publication in the Society's journal, <strong>Comparative</strong> <strong>Parasitology</strong>l If so,<br />

what specific subject area(s) do you feel most qualified to review?<br />

Signature of Applicant Date<br />

Signature of Sponsor (a member)<br />

Mail the completed application along with a check (on a U.S. bank) or money order (in U.S.<br />

currency) for the first year's dues (US$25 for domestic active members and US$28 for foreign<br />

active members) to the Corresponding Secretary-Treasurer, Nancy D. Pacheco, 9708 DePaul Drive,<br />

Bethesda, MD, U.S.A. 20817.<br />

If you wish to pay by credit card (Visa and MasterCard only), please complete the following:<br />

Credit Card: Visa® Q MasterCard® H Expiration Date:<br />

Credit Card Number: O CH d dl I II II I<br />

Signature of Card Holder:<br />

Printed Name Exactly as <strong>State</strong>d on Card:<br />

Date<br />

Amount Charged: US$<br />

Helminthological Society of Washington Home Page: http://www.gettysburg.edu/shendrix/helmsoc.html<br />

Copyright © 2011, The Helminthological Society of Washington


272 COMPARATIVE PARASITOLOGY, <strong>67</strong>(2), JULY <strong>2000</strong><br />

THE HELMINTHOLOGICAL SOCIETY OF WASHINGTON<br />

MISSION AND VISION STATEMENTS<br />

May 7, 1999<br />

THE MISSION<br />

The Helminthological Society of Washington, the prototype scientific organization for parasitological<br />

research in North America, was founded in 1910 by a devoted group of parasitologists in<br />

Washington, D.C. Forging a niche in national and international parasitology over the past century,<br />

the Society focuses on comparative research, emphasizing taxonomy, systematics, ecology, biogeography,<br />

and faunal survey and inventory within a morphological and molecular foundation. Interdisciplinary<br />

and crosscutting, comparative parasitology links contemporary biodiversity studies with<br />

historical approaches to biogeography, ecology, and coevolution within a cohesive framework.<br />

Through its 5 meetings in the Washington area annually, and via the peer-reviewed <strong>Comparative</strong><br />

<strong>Parasitology</strong> (continuing the Journal of the Helminthological Society of Washington in its <strong>67</strong>th<br />

Volume), the Society actively supports and builds recognition for modern parasitological research.<br />

Taxonomic diversity represented in the pages of the Society's journal treats the rich helminth faunas<br />

in terrestrial and aquatic plants, invertebrates, and vertebrates, as well as parasitic protozoans and<br />

arthropods. <strong>Parasitology</strong>, among the most integrative of the biological sciences, provides data critical<br />

to elucidation of general patterns of global biodiversity.<br />

THE VISION<br />

The Helminthological Society of Washington celebrates a century of tradition and excellence<br />

in global parasitology, solving challenges and responding to opportunities for the future of society<br />

and the environment.<br />

Members of the Helminthological Society of Washington contribute to understanding and protecting<br />

human health, agriculture, and the biosphere through comparative research emphasizing<br />

taxonomy, systematics, ecology, biogeography, and biodiversity assessment of all parasites. The<br />

Society projects the exceptional relevance of its programs to broader research and education in<br />

global biodiversity and conservation biology through the activities of its members and its journal,<br />

<strong>Comparative</strong> <strong>Parasitology</strong>.<br />

Copyright © 2011, The Helminthological Society of Washington


*Edna M. Buhrer<br />

* Mildred A. Doss<br />

* Allen Mclntosh<br />

* Jesse R. Christie<br />

•Gilbert F. Otto<br />

* George R. LaRue<br />

*William W. Cort<br />

* Gerard • Dikmans<br />

* Benjamin Schwartz<br />

*Willard H. Wright<br />

*Aurel O. Foster<br />

*Carlton M. Herman<br />

*May Belle Chitwood<br />

*Elvio H. Sadun<br />

E. J. Lawson Soulsby<br />

David R. Lincicome<br />

Margaret A. Stirewalt<br />

•Leo A. Jachowski, Jr.<br />

* Horace W. Stunkard<br />

•Kenneth C. Kates<br />

*Everett E. Wchr<br />

*George R. LaRue<br />

* Vladimir S. Ershov<br />

•Norman R. Stoll<br />

•"Horace W. Stunkard<br />

•Justus F. Mueller<br />

John F. A. Sprent<br />

Bernard Bezubik<br />

Hugh M. Gordon<br />

•W. E. Chambers<br />

*Nathan A. Cobb<br />

* Howard Crawley<br />

*Winthrop D. Foster<br />

•Maurice C. Hall<br />

•Albert Hassall<br />

•Charles W. Stiles<br />

•Paul Bartsch<br />

•Henry E. Ewing<br />

•William W. Cort<br />

•Gerard Dikmans<br />

•Jesse R. Christie<br />

•Gotthold Steiner<br />

•EmmettW. Price<br />

•Eloise B. Cram<br />

•Gerald Thome<br />

•Allen Mclntosh<br />

•Edna M. Buhrer<br />

•Benjamin G. Chitwood<br />

•Aurel O. Foster<br />

•Gilbert F. Otto<br />

•Theodor von Brand<br />

•May Belle Chitwood<br />

•Carlton M. Herman<br />

Lloyd E. Rozeboom<br />

•Albert L. Taylor<br />

David R. Lincicome<br />

Margaret A. .Stirewalt<br />

•Willard H. Wright<br />

•Benjamin Schwartz<br />

•Mildred A. Doss<br />

* Deceased.<br />

ANNIVERSARY AWARD RECIPIENTS<br />

1960<br />

1961<br />

1962<br />

1964<br />

1965<br />

1966<br />

1966<br />

19<strong>67</strong><br />

1969<br />

1969<br />

1970<br />

1971<br />

1972<br />

1973<br />

1974<br />

1975<br />

1975<br />

1976<br />

1977<br />

1978<br />

1979<br />

HONORARY MEMBERS<br />

1959<br />

1962<br />

1976<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

•Philip E. Garrison<br />

*Joseph Goldberger<br />

•Henry W. Graybill<br />

1931<br />

1931<br />

1931<br />

1937<br />

1945<br />

1952<br />

1953<br />

1956<br />

1956<br />

1956<br />

1956<br />

1961<br />

1963<br />

1963<br />

1968<br />

1972<br />

1972<br />

1975<br />

1975<br />

1975<br />

1975<br />

1975<br />

1976<br />

1976<br />

1976<br />

1976<br />

1977<br />

*O. Wilford Olsen<br />

*Frank D. Enzie<br />

Lloyd E. Rozeboom<br />

*Leon Jacobs<br />

Harley G. Sheffield<br />

A. Morgan Golden<br />

Louis S. Diamond<br />

•Everett L. Schiller<br />

Milford N. Lunde<br />

J. Ralph Lichtenfels<br />

A. James Haley<br />

*Francis G. Tromba<br />

Thomas K. Sawyer<br />

Ralph P. Eckerlin<br />

Willis A. Reid, Jr.<br />

Gerhard A. Schad<br />

Franklin A. Neva<br />

Burton Y. Endo<br />

Sherman S. Hendrix<br />

Frank W. Douvres<br />

E. J. Lawson Soulsby<br />

Roy C. Anderson<br />

Louis Euzet<br />

John C. Holmes<br />

Purnomo<br />

Naftale Katz<br />

"Robert Traub<br />

•Alan F. Bird<br />

•Maurice C. Hall<br />

•Albert Hassall<br />

•George F. Leonard<br />

•Everett E. Wehr<br />

•Marion M. Farr<br />

•John T. Lucker, Jr.<br />

George W. Luttermoser<br />

•John S. Andrews<br />

•Leo A. Jachowski, Jr.<br />

•Kenneth C. Kates<br />

•Francis G. Tromba<br />

A. James Haley<br />

•Leon Jacobs<br />

•Paul C. Beaver<br />

•Raymond M. Cable<br />

Harry Herlich<br />

Glenn L. Hoffman<br />

Robert E. Kuntz<br />

Raymond V Rebois<br />

Frank W. Douvres<br />

A. Morgan Golden<br />

Thomas K. Sawyer<br />

*J. Allen Scott<br />

Judith H. Shaw<br />

Milford N. Lunde<br />

•Everett L. Schiller<br />

Harley G. Sheffield<br />

Louis S. Diamond<br />

Mary Hanson Pritchard<br />

Copyright © 2011, The Helminthological Society of Washington<br />

1990<br />

1991<br />

1992<br />

1993<br />

1994<br />

1995<br />

1996<br />

1997<br />

•Charles A. Pfender<br />

•Brayton H.-Ransom<br />

•Charles W. Stiles<br />

1977<br />

1979<br />

1979<br />

1979<br />

1980<br />

1981<br />

1981<br />

1983<br />

1984<br />

1985<br />

1986<br />

1986<br />

1987<br />

1988<br />

1988<br />

1988<br />

1989<br />

1989<br />

1989<br />

1990<br />

1990<br />

1991<br />

1991<br />

1991<br />

1994<br />

1994


VOLUME <strong>67</strong><br />

JULY <strong>2000</strong><br />

CONTENTS<br />

(Continuedfrom Front Cover)<br />

NUMBER 2<br />

RESEARCH NOTES<br />

CANARIS, A. G., AND J. M. KINSELLA. Helminth Parasites of Six Species of Shorebirds (Charadrii) from<br />

Bristol Bay, Alaska, U.S.A. . 250<br />

JEPPS, S. K, AND T. M. GOATER. Colobomatus embiotocae (Copepoda: Philichthyidae) from Shiner Perch,<br />

Cymatogaster aggregate (Osteichthyes: •Embiotocidae) in Canadian Waters 253<br />

CREEL, T. L., G. W. FOSTER, D. J. FORRESTER. Parasites of the Green Treefrog, Hyla cinerea, from Orange<br />

Lake, Alachua County, Florida, U.S.A 255<br />

BAIR, H. D., E. T. LYONS, T. W. SWERCZEK, AND S. C. TOLLIVER. Atypical Specimens of Helminth Parasites<br />

(Anoplocephala perfoliata and Thelazia lacrymalis) of Horses in Kentucky, U.S.A 258<br />

ANNOUNCEMENTS AND NOTICES<br />

OBITUARY NOTICES -. . 164, 168, 180<br />

MUSEUMS FOR DEPOSITING OF SPECIMENS 189<br />

NEW BOOK AVAILABLE 201<br />

MEETING NOTICES , 209<br />

EDITORS' ACKNOWLEDGMENTS . : . 223<br />

HELMINTHOLOGICAL SOCIETY OF WASHINGTON MEETING SCHEDULE „. 240<br />

REPORT OF THE BRAYTON H. RANSOM MEMORIAL TRUST FUND 249<br />

PRESENTATION OF THE 1999 ANNIVERSARY AWARD 261<br />

MINUTES OF MEETINGS OF THE HELMINTHOLOGICAL SOCIETY OF WASHINGTON . 263<br />

AUTHOR INDEX , 265<br />

KEY WORD AND SUBJECT INDEX , . . . 265<br />

MEMBERSHIP APPLICATION 271<br />

MISSION AND VISION STATEMENT OF THE HELMINTHOLOGICAL SOCIETY OF WASHINGTON _.. 272<br />

Date of publication, 24 July <strong>2000</strong><br />

* * *<br />

PRINTED BY ALLEN PRESS, INC., LAWRENCE, KANSAS 66044, U.S.A.<br />

Copyright © 2011, The Helminthological Society of Washington

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!