(ISSN 0161-8202) Journal of ARACHNOLOGY PUBLISHED BY THE AMERICAN ARACHNOLOGICAL SOCIETY VOLUME 36 2008 NUMBER 3 THE JOURNAL OF ARACHNOLOG Y EDITOR-IN-CHIEF: James E. Carrel, University of Missouri-Columbia MANAGING EDITOR: Paula Cushing, Denver Museum of Nature & Science SUBJECT EDITORS: Ecology — Suren Toft, University of Aarhus; Systematics — Mark Harvey, Western Austra- lian Museum and Ingi Agnarsson, University of Akron; Behavior — Gail Stratton, University of Mississippi; Mor- phology and Physiology — Jeffrey Shultz, University of Maryland EDITORIAL BOARD: Alan Cady, Miami University (Ohio); Jonathan Coddington, Smithsonian Institution; William Eberhard, Universidad de Costa Rica; Rosemary Gillespie, University of California, Berkeley; Charles Griswold, California Academy of Sciences; Marshal Hedin, San Diego State University; Herbert Levi, Harvard University; Brent Opeil, Virginia Polytechnic Institute & State University; Norman Platnick, American Museum of Natural History; Ann Rypstra, Miami University (Ohio); Paul Selden, University of Kansas; Matthias Schae- fer, Universitet Goettingen (Germany); William Shear, Hampden-Sydney College; Petra Sierwald, Field Mu- seum; I Min Tso, Tunghai University (Taiwan). The Journal of Arachnology (ISSN 0161-8202), a publication devoted to the study of Arachnida, is published three times each year by The American Arachnological Society. Memberships (yearly): Membership is open to all those interested in Arachnida. Subscriptions to The Journal of Arachnology and American Arachnology (the newsletter), and annual meeting notices, are Included with membership in the Society. Regular, $55; Students, $30; Institutional, $125 . Inquiries should be directed to the Membership Secretary (see below). Back Issues: Patricia Miller, P.O. Box 5354, Northwest Mississippi Community College, Senatobia, Mississippi 38668 USA. Telephone: (601) 562-3382. Undelivered Issues: Allen Press, Inc., 810 E. 10th Street, P.O. Box 368, Lawrence, Kansas 66044 USA. THE AMERICAN ARACHNOLOGICAL SOCIETY PRESIDENT: Paula Cushing (2007-2009), Zoology Department, Denver Museum of Nature & Science, Denver, CO 80205-5798 USA. PRESIDENT-ELECT: Rosemary Gillespie (2007-2009), Environmental Science, Policy & Management, Division of Organisms and Environment, University of California, Berkeley, CA 94720-3114 USA. MEMBERSHIP SECRETARY: Jeffrey W. Shultz (appointed), Department of Entomology, University of Maryland, College Park, MD 20742 USA. TREASURER: Karen Cangialosi, Department of Biology, Keene State College, Keene, NH 03435-2001 USA. SECRETARY: Alan Cady, Dept, of Zoology, Miami University, Middletown, Ohio 45042 USA. ARCHIVIST: Lenny Vincent, Fullerton College, Fullerton, California 92634 USA. DIRECTORS: Elizabeth Jakob (2007-2009), Jason Bond (2006-2008), Greta Binford (2007-2009) PAST DIRECTOR AND PARLIAMENTARIAN: H. Don Cameron (appointed), Ann Arbor, Michigan 48105 USA. HONORARY MEMBERS: C.D. Dondale, H.W. Levi, A.F. Millidge. Cover photo: Clitaetra irenae female (Araneare, Nephilidae) from South Africa. Photo by Matjaz Kuntner. Publication date: 7 November 2008 ©This paper meets the requirements of ANSI/NISO Z39.48-1392 (Permanence of Paper). 2008. The Journal of Arachnology 36:487^-90 A redescription of Varacosa apothetica (Wallace) (Araneae, Lyeosldae) Jamfn M. Dreyer: Department of Biology, Hope College Holland, Michigan 49423, USA. E-mail: dreyerj@hope.edu Abstract. Lycosa apothetica Wallace 1947 is redescribed as a member of the genus Varacosa Chamberlin & Ivie 1942 based on genitalic morphology. The species is freshly illustrated, and information is provided as^ interesting characteristics. Keywords: Lycosa, Trochosa, Florida spiders Like many newly described members of the family Lycosidae Sundevall 1833 from the mid-2Glh century, Lycosa apothetica Wallace 1947 was originally described in the genus Lycosa Latreille 1804. Roewer (1955) placed the species in Varacosa Chamberlin & Ivie 1 942 when he elevated this genus from its subgeneric status within Trochosa C.L. Koch 1847, but his reasons for doing so were not made clear. Varacosa apothetica was not part of Brady’s (1980) Trochosa aver a group nor was it treated as a member of Varacosa by Jimenez & Dondale (1988). Platnick (2008) therefore placed V apothetica within Trochosa along with the rest of Roewer’s Varacosa not included by Jimenez & Dondale (1988). Based on Wallace’s illustrations (1947), it seemed likely that the species belonged to Varacosa. Wallace (1947) examined a total of 57 specimens from the Southeastern United States (8 90° and widening into head of spermathe- cae; fertilization ducts appear suspended above darkened structure posterior to spermathecae when viewed from within. Measurements. — Wallace's original measurements ( 1947) for both the male and female are again reported here in Table 1. Distribution and habitat preferences. — Wallace’s records (1947) indicate that this species is found only in the southeastern USA, from Florida and Georgia west to Mississippi (Figure 8). Most specimens have been collected in Gainesville, Florida. Wallace (1947) reports that “Males have been collected only in October, November, December, and February while females have not been taken after May until October.” It is “secretive,” “usually stays close by, or in, the mouth of it's [sic] retreat” and is “usually found in moist situations in pine flatwoods (pond margins, cypress bay margins, etc.), but may be found occasionally in other situation [sic]”( Wallace 1947). Remarks. — Wallace (1947) lists 57 paratypes in his original description. I examined the majority of those, and three additional specimens collected later. This species bears a synapomorphy with Varacosa : the conspicuous anterior curvatures of the transverse piece. 1 therefore support Roewer’s combination: Varacosa apothetica (Wallace 1947), contra Platnick (2008). In V. apothetica these structures are much wider than those of other Varacosa. Of its congeners, the V. apothetica epigynum most closely resemble V. gosiuta (Chamberlin 1908) and V. shenandoa (Chamberlin & Ivie 1942) (Brady 1980). The palp of the male differs from most Table 1. — Features of V. apothetica (Wallace), taken from Wallace (1947). All measurements in millimeters. See text for abbreviations. Dimension V apothetica <5 V. apothetica ? CW 2.5 2.9 CL 3.5 4.0 POQW 1.0 1.4 POQL 0.8 1.0 PMEW 0.4 0.4 PLEW 0.3 0.3 CH 0.6 0.6 PF 1.3 1.4 PP 0.6 0.7 PT 0.6 0.8 PC/PTC 1.0 1.2 FI 2.7 3.9 PI 1.4 1.5 T1 2.3 2.2 Ml 2.3 2.0 T1 1.5 1.5 Total 1 10.2 11.1 F2 2.4 2.7 P2 1.3 1.5 T2 2.0 2.0 M2 2.1 2.0 T2 1.5 1.5 Total 2 9.3 9.7 F3 2.4 2.6 P3 1.2 1.3 T3 1.7 1.8 M3 2.3 2.2 T3 1.3 1.5 Total 3 8.9 9.4 F4 3.1 3.3 P4 1.2 1.6 T4 2.5 2.8 M4 3.3 3.7 T4 1.8 1.9 Total 4 11.9 13.3 other Varacosa, featuring a large palea and relatively small median apophysis similar to that of V. hoffmannae Jimenez & Dondale 1988. One notable feature of this species is the iridescent quality noted on the body of the males. Male specimens exhibited varying levels of iridescence over their bodies, but each was found to have femora that bore this quality. It is not clear if this is an artifact of the long term preservation of these specimens or a true characteristic of the species. ACKNOWLEDGMENTS I wish to thank the following individuals for their contribution to this work Allen R. Brady and Charles D. Dondale for their identification of this project and helpful encouragement; Norman I. Platnick (AMNH) and G.B. Edwards (FSCA) for the loan of specimens; Lou Sorkin (AMNH) for his assistance with material; C.D. Dondale and Thomas L. Bultman for reading an early version of the manuscript; Dr. Volker Framenau, an anonymous reviewer, and Dr. Ingi Agnarsson for their helpful critiques during the review process; Sarah C. Crews for her suggestions in relation to the illustrations; and the Hope College Department of 490 THE JOURNAL OF ARACHNOLOGY Biology, particularly the Chair T.L. Buitman, for support of this research. LITERATURE CITED Brady, A.R. 1980. Nearctic species of the wolf spider genus Trochosa (Araneae: Lycosidae). Psyche 86:167-212. Chamberlin, R.V. & W. Ivie. 1942. A hundred new species of American spiders. Bulletin of the University of Utah 32:1-117. Jimenez, M.L. & C.D. Dondale. 1988. Descripcion de una nueva especie del genero Varacosa de Mexico. Journal of Arachnology 15:171-175. Platnick, NT. 2008. The World Spider Catalog, Version 8.0. The American Museum of Natural History, New York. Online at http://research.amnh.org/entomology/spiders/catalog/. Accessed 1 6 January 2008. Roewer, C.F. 1955. Katalog der Araneen von 1758 bis 1940, bzw. 1954. Volume 2. Institut Royal des Sciences Naturelles de Belgique, Bruxelles. 923 pp. Wallace, H.K. 1947. A new wolf spider from Florida, with notes on other species. Florida Entomologist 30:33-42. Manuscript received 25 November 2007, revised 10 February 2008. 2008. The Journal of Arachnology 36:491-501 Description of Zabius gaucho (Scorpiones, Buthidae), a new species from southern Brazil, with an update about the generic diagnosis Luis E. Acosta: CONICET - Catedra de Diversidad Animal I, Facultad de Ciencias Exactas, Fisicas y Naturales, Universidad Nacional de Cordoba, Avenida Velez Sarsfield 299, X5000JJC Cordoba, Argentina. E-mail: lacosta@com.uncor.edu Denise M. Candido: Laboratorio de Artropodes, Instituto Butantan, AvenidaVital Brasil, 1500, Butanta, 05503-900, Sao Paulo, Brazil Erica H. Buckup: Museu de Ciencias Naturais, Fundaqao Zoobotanica, Rua Dr. Salvador Franqa, 1427, 90690-000, Porto Alegre, Brazil Antonio D. Brescovit: Laboratorio de Artropodes, Instituto Butantan, AvenidaVital Brasil, 1500, Butanta, 05503-900, Sao Paulo, Brazil. E-mail: adbresc@terra.com.br Abstract. This paper provides the description of a new species in the genus Zabius Thorell (Scorpiones, Buthidae), Z. gaucho n. sp., from four localities in the State of Rio Grande do Sul, Brazil. It differs from Zabius fuscus (Thorell 1 877) and Z. birabeni Mello-Leitao 1938 in details of the telson shape, the longitudinal carinae on mesosomal tergites II-VI, and the number of pectinal teeth. The genus was hitherto known only from Argentina, Z. fuscus being a frequent inhabitant of the central Sierras; Z. birabeni , in turn, is probably a rare and non-orophilous scorpion, collected in scattered localities on the monte/chaco ecotone and in northern Patagonia. The presence of a species of Zabius in southern Brazil lends additional support to the generalized distributional track known as “peripampasic track,” which zoogeographically links the central Sierras Pampeanas with ancient mountains in the southern province of Buenos Aires, southeastern Uruguay and southern Brazil. Keywords: Scorpions, Neotropics, Argentina, taxonomy, new records Resumo. Neste trabalho descrevemos uma especie nova do genero Zabius Thorell (Scorpiones, Buthidae), Z. gaucho n.sp., procendente de quatro localidades do Rio Grande do Sul, Brasil. Distingue-se de Zabius fuscus (Thorell 1877) e Z. birabeni Mello-Leitao 1938 por detalhes da morfologia do telson, das cristas longitudinais dos tergitos do mesossoma, e pelo numero de dentes pectineos. O genero era conhecido ate o momento so para a Argentina, sendo Z. fuscus urn escorpiao muito freqiiente na regiao serrana central; Z. birabeni , no entanto, e aparentemente uma especie nao orofila e rara, coletada em localidades dispersas no ecotono monte/chaco (oeste do pais) e no norte da Patagonia. A presem;a de uma especie de Zabius no sul do Brasil representa um apoio adicional ao padrao generalizado de distribui^ao denominado de “track peripampasico”, que vincula zoogeograficamente as Sierras Pampeanas com sistemas orograficos antigos do sul da Provincia de Buenos Aires, do sudeste do Uruguai e do sul do Brasil. The small buthid genus Zabius Thorell 1894 is restricted to Argentina and previously included only two nominal species: Z. fuscus (Thorell 1877) and Z. birabeni Mello-Leitao 1938. The former species is a very common scorpion occurring in orographic systems in central Argentina, while the latter seems to be a rare species, reported from scattered rockless localities in western and northern Patagonia (Abalos 1953; Maury 1979; Acosta 1989, 1993, 1996; Acosta & Rosso de Ferradas 1996; Mattoni & Acosta 1997; Acosta & Maury 1998; Ojanguren Affilastro 2005). Therefore, the discovery of several specimens of a hitherto undescribed species of Zabius in the State of Rio Grande do Sul, Brazil, represents a remarkable novelty. This new species is described below as Zabius gaucho n. sp. In this paper we also provide new records for Z. fuscus and Z. birabeni and discuss some doubtful reports of the former. Abbreviated synonymies are given to include a few references overlooked by or published after Fet & Lowe (2000). Since the generic diagnosis of Zabius available in the literature is brief (e.g., Mello-Leitao 1945; Abalos 1953), we provide a more complete version, both so as to cover the character states peculiar to Z. gaucho n. sp. and to include several characters introduced by recent taxonomists. Zabius is the southernmost buthid genus occurring in South America and also worldwide. The Neotropical region contains relatively few genera of that family, although one of them, Tityus C.L. Koch 1836, is the most speciose in the order (Fet & Lowe 2000). The presence of a member of Zabius in the state of Rio Grande do Sul has interesting biogeographic implications since it adds further evidence supporting the extension of the generalized distribu- tional track known as the “peripampasic track” (Acosta 1989, 1993) into southern Brazil as briefly discussed below. METHODS Descriptions and line drawings were made using a Leica MSS stereomicroscope equipped with drawing tube. Measurements were taken with a graduated ocular and followed guidelines of Stahnke (1970). Photographs were made using a Canon 400D XTI with a 100 mm Macro-Canon lens. Descriptive terms and abbreviations are as follows: carapacial carinae (Stahnke 1970): Am, anterior median; Cm. central median; Pm, posterior median; Cl, central lateral. Carapacial furrows: based on Stahnke (1970), not abbreviated. Mesosomal carinae (adapted from Vachon 1952): Md, median; Sm, submedian; Sl, sublateral. Carinae of metasomal segments (Francke 1977). Segments I- 491 492 THE JOURNAL OF ARACHNOLOGY IV: Di . dorsal lateral; Lsm, lateral supramedian; Lim. lateral inframedian; Vl, ventral lateral; Vsm. ventral submedian. Segment V: Dl. dorsal lateral; Lm. lateral median; Vl. ventral lateral; Vsm. ventral submedian; Vm. ventral median. Chelal carinae (Soleglad & Sissom 2001; partially also Vachon 1952): Dl, digital; D3, dorsal secondary; D4. dorsal marginal; D5, dorsal internal; I, internomedian; V3, ventrointernal; Vl, ventroexternal; E, external secondary; Va, ventral accessory. Carinae on pedipalp femur and patella are given topological terms, to replace the use of “internal” and “external” (actually, structures inside or outside the tegument); correspondences with Stahnke’s (1970) nomenclature in brackets: prodorsal (instead of dorso-interior), retrodorsal (dorso-exterior), pro- ventral (ventro-interior), retroventral (ventro-exterior), dorsal median (dorso-median), ventral median (ventro-median), retro- lateral median (exterior-median), prolateral median (not mentioned in Stahnke 1970). Acronyms of repositories. — CDA: Catedra de Diversidad Animal I, Facultad de Ciencias Exactas, Fisicas y Naturales, Universidad Nacional de Cordoba, Cordoba, Argentina; IBSP: Institute Butantan, Sao Paulo, Brazil; LEA: Collection of Luis E. Acosta, Cordoba, Argentina; MACN: Museo Argentine de Ciencias Naturales “Bernardino Rivadavia”, Buenos Aires, Argentina; MCN: Museu de Ciencias Naturais, Fundagao Zoobotanica do Rio Grande do Sul, Porto Alegre, Brazil; MNRJ: Museu Nacional, Universidade Federal do Rio de Janeiro, Rio de Janeiro, Brazil; ZMH: Zoologisches Museum, Hamburg, Germany. Additional material examined. Tityus argent inns: ARGEN- TINA: Jujuy: Parque Nacional Calilegua, Toma de Agua (1340 m), 23°41'S, 64°52'W, 25 February 1997 (L. Acosta), 1 = 4.80; P = 0.013). Independent variable b F(l, 44) P Humidity 0.066 1.174 0.284 Temperature 0.485 7.425 0.009 S. S © 60 - ts 50 - ”5 .2 40 - > © 30 - © 20 - a S 3 10 - Z 0 - b q Day Night ^ ^ Day Night Summer Winter J Figure 6. — Comparison of the number of behavioral acts (exclud- ing resting) accomplished per hour by individuals of the harvestman Neosadocus maximus during day and night in two seasons (summer and winter). The horizontal lines represent the mean, the boxes represent the standard deviation, and the vertical lines represent the range (minimum and maximum values in each sample). Different letters above the box-plots indicate significant difference with P < 0.05. N. maximus is greater than those recorded for the harvestmen Discocyrtus oliverioi (25 acts, Elpino-Campos et al. 2001) and Mischonyx cuspidatus (20 acts, Pereira et al. 2004). This difference may be attributed to two main factors: (1) some behavioral acts described in the previous studies were split into two or more different acts in our study (e.g., acts 8-1 1); (2) N. maximus , in fact, presents a number of distinctive behavioral acts that were not recorded before, such as those related to self-grooming (number 16 and 18-20) and social interactions (number 26-28). There was also a marked difference among the three harvestman species in the frequency of the behavioral categories exploration and resting. In D. oliverioi and M. cuspidatus , exploration was the most common behavioral category, comprising nearly 70% of the observations, whereas in N. maximus this category accounted for 28% in winter and 34% in summer. Resting, which ranged from 1 1 to 17% in D. oliverioi and M. cuspidatus, was the most frequent behavioral category for N. maximus comprising 44 to 62% of the observations. These discrepancies could be related to differ- ences in the sampling methods, since Elpino-Campos et al. (2001) and Pereira et al. (2004) concentrated the quantitative observations at night (when the individuals were more active), whereas in the present study we scattered the behavioral samplings equally throughout the whole day. In fact, if we analyze the data obtained only in the night samples for N. maximus, the frequency of resting drops to 22% and of exploration increases to 52%. Therefore, we believe that the data presented in this study are a more realistic scenario of time allocation in harvestmen during the whole day. Contrary to D. oliverioi and M. cuspidatus , individuals of N. maximus did not copulate or lay eggs in captivity, so that the behavioral category reproduction, which comprised one to seven behavioral acts and accounted for 0.1 to 1.8% of the observations in the previous studies, was not included here. Field observations indicate that N. maximus oviposits on the undersurface of leaves and eggs are attended by females (G. Machado unpubl. data). It is possible that the lack of 524 THE JOURNAL OF ARACHNOLOGY appropriate oviposition sites in the terrarium have constrained the reproductive activity of the females. Males, on the other hand, accomplished certain behavioral acts, such as intra- sexual aggressions, that are possibly related to territorial defense and reproduction. Male-male fights have already been described for several harvestman species (e.g., Pabst 1953; Parisot 1962; Edgar 1971; Mora 1990; Macias-Ordonez 2000; Willemart et al. 2006), but this is the first record of this kind of agonistic behavior among gonyleptids. The use of legs IV, which bear large spines and tubercles, shed light on the behavioral roles of the leg armature in N. maximus and other gonyleptid harvestmen as well. Until now, the only function attributed to the armature of legs IV in males was to deliver a nipping upon manipulation, which has been interpreted as a defensive behavior against potential predators (Bristowe 1925; Capocasale & Bruno-Trezza 1964; Gnaspini & Cavalheiro 1998; Machado 2002). A more detailed description of the fights and the functional morphology of legs IV in N. maximus will be described elsewhere (Willemart et al. unpubl. data). Sexual and temporal variation. Males and females of N. maximus differed in the relative frequency of the behavioral categories, with males exploring more frequently than females. This result contrasts with those obtained for D. oliverioi and M. cuspidatus , in which females fed and explored more frequently than males (Elpino-Campos et al. 2001; Pereira et al. 2004). Since both D. oliverioi and M. cuspidatus reproduced in captivity, the higher frequency of feeding activities in females compared to males was attributed to the accumulation of energy for egg production and maturation (Pereira et al. 2004). The higher frequency of exploration by males compared to females in N. maximus may be explained by at least two non-exclusive factors: (1) in this species males seem to be territorial (G. Machado unpubl. data) and thus need to invest time patrolling and exploring their territories; (2) females did not oviposit in captivity and, thus, were not continuously producing eggs, which would reduce the demand for food resources and consequently decrease the frequency of activities related to foraging. Since self-grooming occurs more frequent- ly when individuals are moving around and after feeding (Pereira et al. 2004), the fact that males explored more than females may account for the higher frequency of grooming compared to females. The daily activity of the individuals of N. maximus in captivity was predominantly nocturnal, and light seems to be the most important Zeitgeber promoting synchronization of the activity rhythm. This observation is congruent with data previously obtained in the field, where 100% of the feeding observations in this species occurred at night (Machado & Pizo 2001). The activity pattern, however, presents seasonal variations with a clear change in the phase angle between activity and sunset/sunrise hours from winter to summer. During winter, individuals left the shelters earlier, one hour before dusk. The peak of activity occurred nearly at 20:00 h, and at 02:00 h there was a marked decrease in the frequency of behavioral acts not related to resting. In contrast, during summer, individuals left the shelter only after dusk, remained active throughout the night, and returned to the shelter at about 05:00 h, nearly one hour before the onset of light. In some aspects, this pattern of activity is similar to that described for the cavernicolous harvestman Goniosoma spelaeum, in which the individuals also left the cave earlier in winter when compared to summer (Gnaspini et al. 2003). However, contrary to the present study, individuals of G. spelaeum returned later to the shelter during winter. The authors attributed this change in the phase angle to the time available to forage outside the cave, which is shorter during summer. It is possible that captive individuals of N. maximus can find food faster than individuals of G. spelaeum in the wild and thus can return earlier to their shelters. Another possibility is that different species respond differently to seasonal variations in biotic and abiotic factors. We suggest that future studies investigate how hunger and other physiological constraints modulate long-term changes in the biological rhythms of harvestmen. In our study we demonstrate that temperature, but not humidity, has a positive relationship with the activity of captive individuals of N. maximus, which may explain why the individuals were more active during summer. A similar result has been reported by Capocasale & Bruno-Trezza (1964), who reared individuals of the gonyleptid Acanthopachylus aculeatus (Kirby 1818) in the laboratory and showed that foraging activity seems to be directly related to temperature. However, the predominant nocturnal activity of the species in the field and in the laboratory should not be attributed to temperature since the latter decreases at night. We believe that the decrease in light intensity at dusk, rather than the decrease in temperature, controls the beginning of activity in N. maximus. After darkness, however, air temperature can be an important abiotic factor determining harvestmen activity; the warmer the night the more active are the individuals. Additionally, it is important to stress the role of phylogeny in determining the activity pattern of harvestmen. N. maximus belongs to a clade in which most species are mainly nocturnal. Predominant diurnal activity among gonyleptids seems to be restricted to the clade composed of the subfamilies Progonyleptoidellinae + Caelopyginae (Hoenen & Gnaspini 1999). Methodological approach. — The density of harvestmen used in our terrarium (18 individuals in 0.36 nr) is certainly much higher than the density in the wild, which is no more than 0.04 individuals/nr (G. Machado unpub. data). Crowding may increase the frequency of some behavioral acts, mainly those grouped in the behavioral category “social interactions”. However, for the great majority of the behavioral acts (including those grouped in the behavioral categories “feeding,” “explo- ration,” “self-grooming,” and “resting”) the density of individ- uals in captivity probably had no evident effect. Moreover, since these last four behavioral categories comprise nearly 98% of the behavioral acts in both seasons, we believe that the results obtained under captive conditions for N. maximus provide a realistic scenario of time allocation throughout the day and also in different seasons. The lack of appropriate oviposition sites in our terrarium probably inhibited some behavioral acts, mainly those related to reproduction. We acknowledge this drawback of our laboratory work and recommend that future studies try to reproduce as good as possible the oviposition sites of the study species in the rearing terrarium. Despite some minor problems mentioned above, harvest- men are very convenient animals to keep in captivity since many species are relatively easy to maintain and may live for several months or even years (Willemart 2007). In our case, to OSSES ET AL. — ACTIVITY PATTERN OF NEOSADOCUS MAXIMUS 525 study animals in the laboratory provided the opportunity to quantify the behavior of a great number of individuals in each sampling section and also made it possible to compare the behavioral schedule of the very same subjects in two seasons, which would be very difficult in the wild. Simulating natural variations in climatic conditions, we showed that there were differences in the frequency of five behavioral categories recorded for N. maximus, both at the scale of the day and of the year (winter vs summer). This temporal variation may be endogenously regulated and/or dependent on the environmen- tal variables we manipulated, but we can not actually differentiate between these two possibilities. In this study we also proposed a new sampling procedure for quantitative ethograms that minimizes problems of pseudoreplication. Traditionally, behavioral samplings in quantitative ethograms are accomplished at regular intervals and all subjects are recorded at each scan. Using this procedure, one will face at least two situations that violate the assumption of independent observations required by many statistical tests, especially the frequently used chi-square (see discussion in Kramer & Schmidhammer 1992): (1) one behavioral act may influence the chance of another behavioral act being accomplished in sequence; (2) if the individual A is interacting with the individual B, this behavioral act will be counted twice because the interaction is reciprocal. Yet the “traditional” ethograms have an advantage: the amount of information in each sampling section is high; more specifically n X s, where n is the number of individuals and s is the number of scans. Using the method proposed here, the amount of data in each sampling section is only s because only one individual is scanned at a time. Sampling one individual per scan, however, is exactly the solution for the situation (2) above. Additionally, the median number of times the same individual was scanned per sampling section along the entire period of our study was three. These repeated samplings on the same individuals were spaced out in time, which attenuates the problem exposed in the situation (1). Finally, we also took care of spreading the sampling sections along the time, spacing them with intervals of at least 24 h in order to attenuate the possible influence the activities accomplished in one day could have on activities accomplished in the following day. This is not a standard procedure in “traditional” ethograms, which concentrate the samplings in the periods of more activity or in fixed times of the day. We are aware that sampling the same individuals (in our case 18 harvestmen) is, per se, a source of pseudoreplication, but changing each animal for another one after it was sampled does not seem a reasonable procedure in behavioral studies and would demand a huge quantity of animals. Anyone interested in controlling potential differences among individ- uals may include each one of them as additional factors in the log-linear analysis. Here we avoided this approach because our main goal was to detect general patterns. Conclusions. — In conclusion, we demonstrated that the activity pattern of the neotropical harvestman N. maximus, determined here by the frequency of five behavioral categories, shows sexual and temporal variations. These variations are both quantitative and qualitative since some behavioral acts are restricted to one sex or period of the day. The sampling protocol proposed here should be used in future studies dealing with behavioral repertoires and ethograms because it minimizes the problems of pseudoreplication and provides a more realistic view of the allocation of time and energy for different activities. The great advantage of this method is that it provides a suitable sampling design that generates indepen- dent data, instead of trying to correct the problem of non- independence a posteriori using complicated statistical proce- dures. Our protocol should be useful as a standard method in behavioral samplings not only for harvestman, but for the study of any arthropod reared in captivity. ACKNOWLEDGMENTS We are grateful to Gustavo S. Requena and Bruno A. Buzatto for collecting some individuals used in this study, to Adriano B. Kury for helping with the identification of the studied species, to Paulo De Marco Jr. and Rafael Lourenqo for helping with the statistical analyses, and to Pedro Gnaspini, Sonia Hoenen, Rodrigo H. Willemart, Bruno A. Buzatto, Paulo Enrique C. Peixoto, Gail Stratton, and two anonymous reviewers for insightful comments on an early draft of the manuscript. The authors are supported by grants from CAPES (FO and TMN) and Fundaqao de Amparo a Pesquisa do Estado de Sao Paulo (GM 02/00381-0). LITERATURE CITED Alderweireldt, M. 1994. Day/night activity rhythms of spiders occurring in crop-rotated fields. European Journal of Soil Biology 30:55-61. Aschoff, J. 1960. Exogenous and endogenous components in circadian rhythms. Pp. 11-28. In Biological Clocks. Cold Spring Harbor Symposia on Quantitative Biology, Volume 25. Long Island Biological Association. New York. Bishop, S.C. 1950. The life of a harvestman. Nature Magazine 43:264-267, 276. Bristowe, W.S. 1925. Notes on the habits of insects and spiders in Brazil. Transactions of the Royal Entomological Society of London 1924:475-504. Capocasale, R. & L. Bruno-Trezza. 1964. Biologia de Acanthopachy- lus aculeatus (Kirby, 1819), (Opiliones; Pachylinae). Revista de la Sociedad Uruguaya de Entomologla 6:9-32. Castanho, L.M. & R. Pinto-da-Rocha. 2005. Harvestmen (Opiliones: Gonyleptidae) predating on treefrogs (Anura: Hylidae). Revista Iberica de Aracnologia 1 1:43-45. Cloudsley-Thompson, J.L. 1978. Biological clocks in Arachnida. Bulletin of the British Arachnological Society 4:184-191. Christensen. R. 1997. Log-Linear Models and Logistic Regression. Second edition. Springer-Verlag Inc. New York, New York. 508 pp. Edgar, A.L. 1971. Studies on the biology and ecology of Michigan Phalangida (Opiliones). Miscellaneous Publications, Museum of Zoology, University of Michigan, Ann Arbor 144:1-64. Elpino-Campos, A., W. Pereira, K. Del-Claro & G. Machado. 2001. Behavioural repertory and notes on natural history of the neotropical harvestman Discocyrtus oliverioi (Opiliones: Gonylep- tidae). Bulletin of the British Arachnological Society 12:144-150. Foelix, R.F. 1996. Biology of Spiders. Second edition. Oxford University Press, New York. 330 pp. Friebe, B. & J. Adis. 1983. Entwicklungzyklen von Opiliones (Arachnida) im Schwarzwasser-Uberschwemmungswald (Igapo) des Rio Taruma Mirim (Zentralamazonien, Brasilien). Amazoni- ana 8:101-1 10. Gnaspini, P. 1996. Population ecology of Goniosonia spelaeum, a cavernicolous harvestman from southeastern Brazil (Arachnida: Opiliones: Gonyleptidae). Journal of Zoology 28:137-151. 526 THE JOURNAL OF ARACHNOLOGY Gnaspini, P. & A.J. Cavallieiro. 1998. Chemical and behavioral defenses of a Neotropical cavernicolous harvestman Goniosoma spelaeum (Opiliones, Laniatores, Gonyleptidae). Journal of Ara- chnology 26:81-90. Gnaspini, P., F.H. Santos & S. Hoenen. 2003. The occurrence of different phase angles between contrasting seasons in the activity patterns of the cave harvestman Goniosoma spelaeum (Arachnida, Opiliones). Biological Rhythm Research 34:31-49. Hoenen, S. & P. Gnaspini. 1999. Activity rhythms and behavioral characterization of two epigean and one cavernicolous harvestmen (Arachnida, Opiliones, Gonyleptidae). Journal of Arachnology 27:159-164. Kramer, M. & J. Schmidhammer. 1992. The chi-squared statistic in ethology: use and misuse. Animal Behaviour 44:833-841. Kury, A.B. 2003. Annotated catalogue of the Laniatores of the New World (Arachnida, Opiliones). Revista Iberica de Aracnologia, vol. especial monografico 1:1-337. Lehner, P.N. 1996. Handbook of Ethological Methods. Garland STPM Press, New York. 672 pp. Machado, G. 2002. Maternal care, defensive behavior, and sociality in Neotropical Goniosoma harvestmen (Arachnida, Opiliones). Insectes Sociaux 49:1-6. Machado, G. & P.S. Oliveira. 1998. Reproductive biology of the Neotropical harvestman Goniosoma longipes (Arachnida, Opi- liones: Gonyleptidae): mating and oviposition behaviour, brood mortality, and parental care. Journal of Zoology 246:359-367. Machado, G. & M.A. Pizo. 2000. The use of fruits by the Neotropical harvestman Neosadocus variabilis (Opiliones, Laniatores, Gony- leptidae). Journal of Arachnology 28:357-360. Machado, G., R.L.G. Raimundo & P.S. Oliveira. 2000. Daily activity schedule, gregariousness, and defensive behavior in the Neotrop- ical harvestman Goniosoma longipes (Arachnida: Opiliones: Gony- leptidae). Journal of Natural History 34:587-596. Machado, G. & D.M. Vital. 2001. On the occurrence of epizoic algae and liverworts on the harvestmen Neosadocus aff. variabilis (Opiliones: Gonyleptidae). Biotropica 33:535-538. Macias-Ordonez, R. 2000. Touchy harvestmen. Natural History 109:58-61. Marc, P. 1990. Nycthemeral activity rhythm of adult Clubiona corticalis (Walckenaer, 1802) (Araneae, Clubionidae). Acta Zool- ogica Fennica 190:279-285. Martens, J. 1993. Bodenlebende Arthropoda im zentralen Himalaya: Bestansaufnahme, Wege zur Viefalt und Okologische Nischen. Pp. 23-250. In Neue Forschungen im Himalaya. (U. Scheinfurth, ed.). Erdkundliches Wissen. Volume 112. Franz Steiner Verlag, Stuttgart. Martin, P. & P. Bateson. 1993. Measuring Behaviour: an Introduc- tory Guide. Second edition. Cambridge University Press, New York. 222 pp. Mora, G. 1990. Parental care in a Neotropical harvestman, Zygopachylus albomarginis (Arachnida, Opiliones: Gonyleptidae). Animal Behaviour 39:582-593. Pabst, W. 1953. Zur Biologie der mitteleuropaischen Troguliden. Zoologische Jahrbucher, Abteilung fiir Systematik, Okologie und Geographic der Tiere 82:1-156. Parisot, C. 1962. Etude de quelques opilions de Lorraine. Vie et Millieu 13:179-197. Pereira, W., A. Elpino-Campos, K. Del-Claro & G. Machado. 2004. Behavioral repertory of the Neotropical harvestman Ilhaia cuspidata (Opiliones, Gonyleptidae). Journal of Arachnology 32:22-30. Pfeifer, H. 1956. Zur Okologie und larvalsysteinatik der Weber- knechte. Mitteilungen aus dem Zoologischen Museum in Berlin 32:59-104. Phillipson, J. 1959. The seasonal occurrence, life histories and fecundity of harvest-spiders (Phalangida, Arachnida) in neighbor- hood of Durham City. Entomologist's Monthly Magazine 95:134-138. Polis, G.A. 1980. Seasonal patterns and age-specific variation in the surface activity of a population of desert scorpions in relation to environmental factors. Journal of Animal Ecology 49:1-18. Polis, G.A. & R.D. Farley. 1979a. Behavior and ecology of mating in the cannibalistic scorpion, Paruroctonus mesaensis Stahnke (Scor- pionida: Vaejovidae). Journal of Arachnology 7:3-46. Polis, G.A. & R.D. Farley. 1979b. Characteristics and environmental determinants of natality, growth and maturity in a natural population of the desert scorpion Paruroctonus mesaensis (Scor- pionida: Vaejovidae). Journal of Zoology 187:517-542. Punzo, F. 1998. The Biology of Camel Spiders (Arachnida, Solifugae). Kluwer Academic Publishers, Dordrecht, The Nether- lands. 301 pp. Todd, V. 1949. The habits and ecology of the British harvestmen (Arachnida, Opiliones), with special reference to those of the Oxford District. Journal of Animal Ecology 18:209-229. Tsurusaki. N. 2003. Phenology and biology of harvestmen in and near Sapporo, Hokkaido, Japan, with some taxonomical notes on Nelima suzukii n.sp. and allies (Arachnida: Opiliones). Acta Arachnologica 52:5-24. Warburg, M.R. & G.A. Polis. 1990. Behavioral responses, rhythms, and activity patterns. Pp. 224-246. In The Biology of Scorpions. (G.A. Polis, ed.). Stanford University Press, Stanford, California. Willemart, R.H. 2001. Egg covering behavior of the neotropical harvestman Promitobates ornatus (Opiliones, Gonyleptidae). Jour- nal of Arachnology 28:249-252. Willemart, R.H. 2002. Cases of intra- and inter-specific food competition among Brazilian harvestmen in captivity (Opiliones, Laniatores, Gonyleptidae). Revue Arachnologique 14:49-58. Willemart, R.H. 2007. Rearing and maintenance of harvestmen in captivity. Pp. 520-524. In Harvestmen: The Biology of Opiliones. (R. Pinto-da-Rocha, G. Machado & G. Giribet, eds.). Harvard University Press, Cambridge, Massachusetts. Willemart. R.H., M.C. Chelini, R. de Andrade & P. Gnaspini. 2007. An ethological approach to a SEM survey on sensory structures and tegumental gland openings of two neotropical harvestmen (Arachnida, Opiliones, Gonyleptidae). Italian Journal of Zoology 74:39-54. Willemart, R.H.. J.P. Farine, A.V. Peretti & P. Gnaspini. 2006. Behavioral roles of the sexually dimorphic structures in the male harvestman Phalangium opilio (Opiliones. Phalangiidae). Canadian Journal of Zoology 84:1763-1774. Willemart, R.H. & P. Gnaspini. 2003. Comparative density of hair sensilla on the legs of cavernicolous and epigean harvestmen (Arachnida: Opiliones). Zoologisher Anzeiger 242:353-366. Willemart, R.H. & P. Gnaspini. 2004a. Spatial distribution, displacement, gregariousness and defensive behavior in the Brazilian cave harvestman Goniosoma albiscriptum (Arachnida, Opiliones, Laniatores). Animal Biology 54:221 235. Willemart, R.H. & P. Gnaspini. 2004b. Breeding biology of the cavernicolous harvestman Goniosoma albiscriptum (Arachnida, Opiliones, Laniatores): sites of oviposition, egg-batches, charac- teristics and subsocial behaviour. Invertebrate Reproduction and Development 45:15-28. Williams, G.C. 1962. Seasonal and diurnal activity of harvestmen (Phalangida) and spiders (Araneida) in contrasted habitats. Journal of Animal Ecology 31:23-42. Manuscript received 29 August 2006. revised 26 February 2008. 2008. The Journal of Arachnology 36:527-532 First male sperm precedence in multiply-mated females of the cooperative spider Anelosimus studiosus (Araneae, Theridiidae) Thomas C. Jones: Department of Biological Sciences, East Tennessee State University, Johnson City, Tennessee 37614, USA. E-mail: jonestc@etsu.edu Patricia G. Parker: Department of Biology, University of Missouri at St. Louis, St. Louis, Missouri 63121, USA Abstract. Patterns of sperm usage in multiply-mated females have profound fitness consequences for males, and create strong selective pressure on male behavior. In the cooperative theridiid spider Anelosimus studiosus Hentz 1850 adult males are tolerated in females’ webs, and females have been observed to mate multiply with different males. In this experiment, virgin females were mated with two different males on consecutive days under controlled conditions to determine paternity patterns and behavioral responses of males to non-virgin females. The paternity of broods was analyzed using randomly amplified polymorphic DNA (RAPDs). Fifteen broods were analyzed and complete first male sperm precedence was found. Mating behavior differed between first and second males with the first males attempting fewer intromissions, but having a longer total time of intromission. This suggests that the second males are either prevented from normal copulation, or are reacting to the different condition of the females. The sperm precedence pattern is discussed with respect to its ramifications for male behavior, juvenile inclusive fitness, and the evolution of cooperative behavior. Keywords: Social spiders, mating behavior. RAPDs, sexual selection When females mate multiply, sperm precedence patterns can affect the fitness of all individuals involved. For males who have both mated with the same female, the fitness conse- quences are clear and directly related to the proportion of her brood they have fertilized (Trivers 1972). Because of this, males often compete with each other for access to females, or to be chosen as mates by females (Andersson 1994). Males may also compete for fertilizations after copulation through such avenues as mate guarding or copulatory plugs, the necessity or efficacy of which is affected by sperm precedence (Parker 1984). Patterns of sperm precedence may affect the female’s fitness by influencing the genetic variability of her brood, or the proportions of her brood fertilized by males of differing quality. There is also accumulating evidence of females manipulating fertilization patterns of their broods in response to male attributes (Eberhard 1996). Examples of this in spiders include selective sperm storage in response to copulation duration (Bukowski & Christenson 1997a & b), and the fact that, with paired spermathecae, spiders may be pre-adapted for paternity manipulation (Snow & Andrade 2005). Patterns of sperm precedence will also affect the composi- tion of full and half-sibs within broods of polyandrous females. In social species, the relatedness among brood-mates can have profound effects on their fitness (Hamilton 1964). The relatedness among group members is therefore important for a complete understanding of the selective costs and benefits to group living (for review see Caraco & Giraldeau 1991). Mating systems are particularly important to the evolution of sociality in spiders. Social spider colonies are generally inbred (Riechert & Roeloffs 1993; Johannesen et al. 2002). In fact, genetic analysis of Anelosimus eximius Keyserling 1884 colonies suggests that there is no gene Bow at all among colonies (Smith & Hagen 1996). It is likely that the cooperative behaviors and female-biased sex ratios of cooperatively social spider species are maintained by interdemic selection, fostered by the high levels of relatedness among colony members (Aviles 1997). Isolated local populations of asocial or subsocial spiders will become inbred through genetic drift, which could then promote the evolution of cooperative behaviors. The rate at which a population loses genetic diversity (i.e., the effective population size) is affected by its mating system in that monogamous populations lose diversity faster than promiscuous populations (Parker & Waite 1997). This inbreeding is likely an important factor in the evolution- ary transition from subsociality to permanent sociality in spiders (Bilde et al. 2005). Anelosimus studiosus Hentz 1850 is a relatively small (about 8 mm long) theridiid spider which ranges from Argentina to New England (Agnarsson 2006; Agnarsson et al. 2007). This species is common in the southeastern USA and can be found in extremely high densities along waterways (Jones et al. 2007). This species is described as subsocial (Wilson 1971) and, specifically, “prolonged subsocial” (Rayor & Taylor 2006), in that juveniles and adult males are tolerated in an adult female’s web, but other adult females usually are not (Brach 1977; but see Furey 1998; Jones et al. 2007). Previous experiments have demonstrated that, under controlled labo- ratory conditions, colony prey capture increases with the age and number of juveniles in the colony (though resources per individual decline with colony size), and variation in prey mass decreases with the number of juveniles present (Jones & Parker 2000). Females can produce up to at least three broods over their lives usually with several weeks between broods (Jones, unpubl. data). It has also been shown that in semi-natural conditions, delayed juvenile dispersal benefits juvenile survi- vorship and development as well as the mother’s ability to produce future broods (Jones & Parker 2002). While it is clear that individual juveniles are better off in their natal group than on their own, the exact relationship of individual fitness to group size is not yet known. Whatever this relationship, a juvenile’s fitness is likely to be affected by its relatedness to its brood-mates. We have observed that females will mate 527 528 THE JOURNAL OF ARACHNOLOGY multiply if presented with more than one male, so broods may be composed of either full-sibs, or a mixture of full and half- sibs. Members of the family Theridiidae are “entelegyne” spiders, in that the female reproductive tract has a conduit morphology, with sperm leaving the storage organ to fertilize eggs through a different opening than that into which they were deposited (Foelix 1996). It has been suggested that this morphology will put the first mate’s sperm closest to the point of fertilization, and thus lead to first mate sperm precedence (Austad 1984). However, studies of sperm precedence in entelegyne spiders have yielded estimates of proportions of first male from 0.95 to only 0.37 (reviewed by Elgar 1998). In the theridiid Australian redback spider the mean first male precedence was 0.44, but ranged from 0.0 to 1.0 (Andrade 1996). In this study we seek to determine sperm precedence in A. studiosus by sequentially pairing females with two males, recording mating behavior, and using the animals’ DNA and RAPD analysis to determine parentage. METHODS Collection and Rearing. — Anelosimus studiosus colonies containing juveniles were collected in southern Louisiana in September of 1999 from bayous in Tickfaw State Park (30°22'N, 90°37'W) and Fontainebleau State Park (30°20'N, 90 02' W). The bayous were accessed by pirogue and the webs were mostly collected from low-hanging cypress branches (voucher specimens are deposited in the Museum of Biological Diversity at The Ohio State University, Columbus, Ohio). We maintained the colonies in plastic containers (18 X 14 x 6 cm) that were laced with sewing thread to provide substrate for web building. The colonies were kept at buffered ambient temperatures (20-28 C) under natural light conditions, were fed Drosophila melanogaster and D. virilis ad libidum, and misted with distilled water three times a week. We kept the colonies in a greenhouse under natural lighting conditions. In order to ensure the virginity of experimental animals, as the juveniles approached maturity, females were isolated in new containers, and the penultimate males were grouped together (by natal colony) in another set of containers. Experimental Matings. — Females that had undergone their final molt in isolation were selected for this experiment. Twenty-five females from eight different colonies, but no more than four females from any one colony, were used. We released males from two different colonies (other than the colony of their prospective mate) into a female’s container one at a time, on consecutive days. Matings were videotaped with a Sony digital 8 camera for behavioral analysis. The resolution of the video did not allow fine details of palpal insertions, such as the extension of the embolus, to be observed. We estimated copulation as periods when the male’s palps were resting stationary against the female’s genital opening. The onset of copulation was recorded as when the palp would ease against the genital opening and stop, and the end of copulation was recorded when the male pushed against the female and broke loose with a conspicuous jerk. We quantified the number and timing of copulations and copulation attempts. After mating, the males were frozen for later DNA extraction. Mating usually commenced within a few minutes after the introduc- tion of the males. We removed the males after they had broken copulation, moved off from the female, and for ten minutes showed no further attempt at copulation. On four occasions the males had not attempted copulation after 15 min. These males were removed from the container and the process was restarted with new males. Abstinent males were not reused. After the females had mated the second time, we returned them to the greenhouse rearing conditions. Females that produced broods were allowed to rear them through the third instar, after which the female and juveniles were frozen for DNA extraction. DNA Extraction and PCR. — To the 1.5 ml tubes containing the frozen spiders, 200 pi of CTAB and 1.5 pi Proteinase K (100 pg/ml) were added. The spiders were thoroughly ground with a pestle in the tube and incubated at 60° C for 1 h. One extraction with 100 pi of phenol and 100 pi of CIA (24:1 chloroforrmisoamyl alcohol), and one extraction with 200 pi of CIA were performed. The samples were ethanol precipitat- ed, resuspended in 100 pi of distilled water, and stored at 4° C. The concentration of the samples was estimated by gel comparison with concentration standards. Randomly amplified polymorphic DNA (RAPDs) uses single relatively non-specific ten base pair primers (synthesized at OSU) to amplify regions of the genome that contain complementary primer annealing sites. The regions that are amplified are arbitrary but heritable, and therefore, useful (Williams et al. 1990). Under similarly controlled mating conditions, RAPDs were used to assign sperm precedence in a beetle (Carbone & Rivera 2003). The reactants for an individual 14 pi reaction consisted of: 9.9 pi, UV irradiated distilled water; 1.5 pi of 1 pM dNTPs; 1.5 pi reaction buffer (10 mM Tris HC1, pH 8.3; 50 mM KC1; 2 mM MgCl); 0.8 pi of 10 pM primer; 0.1 pi Taq DNA polymerase (5 U/pl); 1.2 pi template DNA (approx. 25 ng/pl). The reactions were run through four initial “touch down” cycles (94° C for 1 min; 35° C for 1 min; 0.3 slope to 72° C for 2 min), and then 32 amplification cycles (94° C for 10 s; 35° C for 30 s; 12° C for 30 s). The finished reactions were held at 4°C until they were visualized. For visualization, the amplified products were run out on a 1.2% agarose gel (80-120 V), stained with ethidium bromide, then visualized and photographed under UV light. Paternity analysis. — RAPDs are dominant markers, and band presence/absence is particularly sensitive to reaction conditions because of the short length of the primers. Therefore, repeatability of RAPD markers has been prob- lematic, making them not as robust in parentage analyses as some other molecular techniques (e.g., microsatellites or mullilocus minisatellite DNA fingerprinting; for review see Parker et al. 1998). In this experiment, however, RAPDs were useful to assess paternity of broods because the pool of potential fathers is limited and known, and because repeatability was confirmed. A unique bands analysis was used to assign the father of each brood member. On the gels, the mother and two potential fathers were run as triads twice, flanking the offspring lanes. Bands that were observed in the lanes of one of the males, but not in the lanes of the other male or mother, were scored for their presence in the offspring lanes. Multiple primers were screened for the families until a total of at least two diagnostic bands were found for each juvenile. Such an analysis is simple and robust since no inference is made from band absence, and the JONES & PARKER SPERM PRECEDENCE IN A COOPERATIVE SPIDER 529 repeatability of each diagnostic band is confirmed by amplifying the triads of adults twice, and running them on the flanking lanes on both sides of the gels. RESULTS Mating behavior. — When the males were placed in the containers near the females, they would typically remain motionless for up to 1 min. They would then begin to move around in the web while rapidly drumming their first pair of legs on the silk. The movement of the males appeared undirected until the females moved within the web, at which point the males would begin to move toward the females while still drumming. As the males approached, the females would typically bounce in the web apparently signaling sexual receptivity because the males would move more quickly toward them afterwards. The males continued drumming even as they made contact with the females. The males would orient themselves to face the same direction as the female, with their ventral surfaces adjacent, but with no consistent absolute orientation. As the males moved into position, the amplitude of their drumming eased to a stop, which was taken to be the onset of copulation. After copulation the spiders separated with a conspicuous jerk, followed by the males moving a short distance from the females (1-2 cm). If only one copulation had taken place, the males would resume drumming and repeat the courtship, but would typically move in more quickly and insert on the other side. The females in this experiment, in all cases, appeared receptive to both males. Also, no occurrences or apparent attempts of sexual cannibalism were observed. As measured by the number of copulation onsets and breaks, first males had fewer copulations (mean 2.2, range 2- 4) than second males (mean 3.7, range 2-12; Mann-Whitney U = 489, P < 0.001, Fig. 1). Only three of the 25 first males had more than two copulations, and in those cases there were one or two short copulations followed by two long ones. The total time spent in copula was longer for first males (mean 44.3 min, range 34.7-52.4) than for second males (mean 15.9 min, range 1.5-42.1; t = 11.3, P < 0.001, Fig. 1). In six cases the second males had more than three copulations, but their total time of copulation (mean = 11.0 min) was significantly shorter than the second males that had three or fewer copulations (mean = 19.6; t = 1.9, P = 0.04). Considering individual females, their first mate’s total time of copulation was not related to their second mate’s total time (Fig. 2). The duration of the first male’s first copulation was a strong predictor of the duration of his second copulation (in cases where there were more than two copulations, the initial apparent “false starts’’ [copula- tions lasting less than one minute] were excluded: Fig. 3). Paternity analysis. — Of the 25 females in the mating experiment, 22 produced egg cases, of which 17 had juveniles emerge (which was considerably lower than the mean of 36 juveniles observed in nature; Brach 1977). The average number of juveniles per family was 11.6 (range 4-21). Two of the families were unusable because one of the males (in each family) could not be amplified by PCR. Of the fifteen remaining families, all 168 of the juveniles were assigned to the first male. While the assignments were based on the presence of at least two of the first male’s unique bands, the lack of the second male’s unique bands in juvenile lanes further confirmed the assignments. Male Male Male Male Figure 1. — Comparison of the mean number of copulations and total duration of copulation between first and second males in Anelosimus studiosus. Reported are means with standard error bars. DISCUSSION Using a direct DNA-based analysis of parentage, we found complete first male sperm precedence in A. studiosus. We also found significant differences in the mating behavior of first and second males. The patterns of sperm precedence, the increased frequency of starting and stopping copulation with less actual time spent in copula of second males, suggests that the second males could not successfully copulate. There is evidence from other spider species that second males can be prevented from successfully copulating. Copu- lation “plugs” made by a hardening of seminal fluids in the female reproductive tract or by the tip of the male’s intromittent organ breaking off, have been reported for several spider species Phidippus johnsoni Peckham & Peckham 1983 (Jackson 1980), Agelena limbata Thorell 1897 (Matsu- moto 1993; Yoward & Oxford 1996; Schneider et al. 2005). However, the efficacy of these plugs in preventing subsequent T otal Time of F irst Male Copulation Figure 2. — Plot of a female’s total time of copulation with the first male versus her total time of copulation with the second male in Anelosimus studiosus. 530 THE JOURNAL OF ARACHNOLOGY q 15 20 25 30 Duration of First Male’s First Copulation (Min) Figure 3. — Plot of the duration of the first male’s first copulation versus the total time of his second copulation in Anelosimus studiosus. fertilizations is mixed among taxa, and they may instead function to increase the sperm retention and survival (reviewed in Huber 2005). In another theridiid spider, fertilization patterns are affected by where the broken organ tip rests in the female reproductive tract (Snow et al. 2006). Across spider taxa, male genital mutilation appears to be an indicator of strong selection on paternity protection, being correlated with the evolution of male sacrifice behavior and size dimorphism (Miller 2007). We did not determine in this study whether or not the second male was in any way prevented from copulation. The fact that four of the second males’ copulatory behavior was similar to the first males’ (having two intromissions totaling > 30 min), suggests that copulatory plugs are not ubiquitous, or at least not completely effective. However, we have observed a second male repeat- edly moving in and breaking from a female, apparently struggling to insert his palps. Interestingly, first male total copulation time was only 16% that observed in a congener (Klein et al. 2005). The fact that second males spent less total time in copulation contradicts previous findings in an araneid spider (Micrathena gracilis Walckenaer 1805) in which second males copulated over twice as long as the first male (Bukowski & Christenson 1997b), but similar to patterns found in a tetragnathid spider ( Tetragnatha versicolor Walckenaer 1842; Danielson-Francois & Bukowski 2004). In the latter case sperm release was equal between first and second males. A similar experiment in which males were introduced to females that had only been mated on one side of their tract suggested that males are responding specifically to the reproductive status of the female’s reproductive tract, rather than the female’s behavior or overall condition (Bukowski et al. 2001). Second males also copulated for shorter periods in a cellar spider, P/iolcus phalagioides Fuesslin 1775 (Schafer & Uhl 2002). In this case, however, the second males had a higher proportion of reproductive success, apparently as a result of their ability to remove the first male’s sperm. Since second males generally attempted copulation, the differences in their mating behavior seems most likely because they are prevented from normal copulation. The possibility remains however, that they could be altering their behavior in response to the previously mated condition of the female. There are examples of non-virgin spiders being less attractive to males. The presence of sex pheromones has been documented across a wide range of spider species and, in some cases, these pheromones are volatile (Shultz & Toft 1993; Miyashita & Hayashi 1996; Rovner 1996; Costa et al. 1997; Searcy et al. 1999), and in other cases are contact based (Trabalon et al. 1997, 1998). In one linyphiid spider (Nereine litigiosa Keys 1886) pheromones are incorporated in the female's web, and males destroy the web prior to mating, reducing the probability that a second male will find her (Watson 1986). Whether pheromones exist in this species is not known, but pheromone-like compounds have been extracted from the cuticle of its congener A. eximius (Bagneres et al. 1997). It is possible that, even though the female remains sexually receptive after mating, her production of pheromones decreases, thus making her less attractive to second males. The speed with which the first and second males begin drumming and searching for the female might give insight into pheromone levels, but such a measure would be confounded by the introduction of the males, which was not standardized in terms of their distance from the female. This study found complete first male sperm precedence, accentuating the question why an A. studiosus female should mate multiply at all. There are many potential costs to mating such as loss of foraging opportunities, increased predation risk, and disease transmission (reviewed in Lewis 1987). Male spiders provide no parental care, and it is unlikely that any substances that males transfer along with sperm provide direct benefits to the females as has been observed in some insects (Gwynne 1984; Boggs 1990). With linyphiid spiders the male cohabitats in the female’s web. eating prey, until he has mated with her; females apparently mate with these males to induce them to leave (Watson 1993). However, there is evidence from the same species that multiple mating has indirect benefits in terms of the size and growth rate of juveniles (Watson 1998). In the pisaurid species Pisaura mirabilis (Clerk 1757), in which females mate with multiple males, males present nuptial prey items (Drengsgaard & Toft 1999). Mating multiply allows the possibility of cryptic female choice in which she chooses the sperm of the male she prefers (Eberhard 1996). Again, given complete first male precedence, this seems unlikely to be occurring with this species. Perhaps the simplest benefit to a female from multiple mating would be to ensure that all her eggs get fertilized. It may also be that there is no selective benefit to mating with multiple males, and that A. studiosus females simply remain receptive from maturity until their abdomens are distended with eggs regardless of the number of times they have mated. Previous work on this species has demonstrated that by delaying dispersal and remaining part of their natal colony, juveniles enhance their survival and development (Jones & Parker 2002). This can be extrapolated to suggest that juveniles’ direct fitness benefits from delayed dispersal. This study finds complete first male sperm precedence within the broods of doubly-mated A. studiosus females. This suggests that if an individual juvenile’s presence in the colony contributes to the survivorship of its brood-mates, its indirect fitness (sensu Hamilton 1964) would be maximized because JONES & PARKER— SPERM PRECEDENCE IN A COOPERATIVE SPIDER 531 they are all full-sibs. In this species it is now documented that there are colonies that contain multiple adult females in North America (Furey 1998), the incidence of which increases with latitude (Jones et al. 2007). If these colonies develop by- non- dispersal of juveniles, fertilization patterns could have profound effects on the genetic structure of these large colonies. Finding complete first-male sperm precedence may be surprising, but these results should be taken with some caution given the highly controlled conditions. Factors such as the number and timing of matings, which may influence precedence (Eberhard 1996), were held constant. Currently studies are underway exploring relatedness within and among natural colonies using microsatellite loci. ACKNOWLEDGMENTS We wish to thank the Department of Evolution, Ecology, and Qrganismal Biology at The Ohio State University. Special thanks to George Keeney and Lisa Wallace for technical assistance. We also thank T. Grubb, E. Marschall, G. Uetz, and members of the Parker lab for useful discussion. Thanks also to G. Stratton and two anonymous reviewers for their insightful comments. LITERATURE CITED Agnarsson, I. 2006. Revision of the new world eximius lineage of Anelosimus and a phylogenetic analysis using worldwide exem- plars. Zoological Journal of the Linnean Society 146:453-593. Agnarsson, L, W.P. Maddison & L. Aviles. 2007. The phylogeny of the social Anelosimus spiders (Araneae: Theridiidae) inferred from six molecular loci and morphology. Molecular Phylogenics and Evolution 43:833-851. Andersson, M. 1994. Sexual Selection. Princeton University Press, Princeton, New Jersey. 624 pp. Andrade, M.C.B. 1996. Sexual selection for male sacrifice in the Australian redback spider. Science 271:70-72. Austad, S.N. 1984. Evolution of sperm priority patterns in spiders. Pp. 233-249. In Sperm Competition and the Evolution of Animal Mating Systems. (R.L. Smith, ed.). Academic Press, London. Aviles, L. 1997. Causes and consequences of cooperation and permanent sociality in spiders. Pp. 476-498. In The Evolution of Social Behavior in Insects and Arachnids. (J.C. Choe & B.J. Crespi, eds.). Cambridge University Press, Cambridge, UK. Bagneres, A.G., M. Trabalon, G.J. Blomquist & S. Schulz. 1997. Waxes of the social spider Anelosimus eximius (Araneae, Theridi- idae): Abundance of novel n-propyl esters of long-chain methyl- branched fatty acids. Archives of Insect Biochemistry and Physiology 36:295-314. Bilde, T., Y. Lubin, D. Smith, J. Schneider & A.A. Maklakov. 2005. Transition to social inbred mating systems: role of inbreeding tolerance in a subsocial predecessor. Evolution 59:160-174. Boggs, C.L. 1990. A general model of the role of male-donated nutrients in female insects’ reproduction. American Naturalist 136:598-617. Brach, V. 1977. Anelosimus studiosus (Araneae: Theridiidae) and the evolution of quasisociality in theridiid spiders. Evolution 31: 154-161. Bukowski, T.C. & T.E. Christenson. 1997a. Natural history and copulatory behavior of the spiny orbweaving spider Micrathena gracilis (Araneae, Araneidae). Journal of Arachnology 25:307-320. Bukowski, T.C. & T.E. Christenson. 1997b. Determinants of sperm release and storage in a spiny orbweaving spider. Animal Behaviour 53:381-395. Bukowski, T.C., C.D. Lynn & T.E. Christenson. 2001. Copulation and sperm release in Gastracantha canceriformis (Araneneae: Arandeidae): differential male behaviour based on female mating history. Animal Behaviour 62:887-895. Caraco, T. & L.-A. Giraldeau. 1991. Social foraging: producing and scrounging in a stochastic environment. Journal of Theoretical Biology 153:559-589. Costa, F.G., C. Viera & G. Francescoli. 1997. Male sexual behavior elicited by a hybrid pheromone: a comparative study on Lycosa thorelli, L. carbonelli, and their progeny (Araneae, Lycosidae). Canadian Journal of Zoology 75:1845-1856. Oanielson-Francois, A.M. & T.C. Bukowski. 2004. Female mating history influences copulation behavior but not sperm release in the orb-weaving spider Tetragnatha versicolor (Araneae, Tetragnathi- dae). Journal of Insect Behavior 18:131-148. Drengsgaard, I.L. & §. Toft. 1999. Sperm competition in a nuptial feeding spider, Pisaura mirabilis. Behaviour 136:877-897. Eberhard, W.G. 1996. Female Control: Sexual Selection by Cryptic Female Choice. Princeton University Press, Princeton, New Jersey. 472 pp. Elgar, M.A. 1998. Sexual selection and sperm competition in arachnids. Pp. 307-337. In Sperm Competition and Sexual Selection. (T.R. Birkhead & A.P. Moller, eds.). Academic Press, London. Foelix, R.F. 1996. Biology of Spiders, Second edition. Oxford University Press, New York. 330 pp. Furey, R.E. 1998. Two cooperatively social populations of the theridiid spider Anelosimus studiosus in a temperate region. Animal Behaviour 55:727-735. Gwynne, D.T. 1984. Courtship feeding increases female reproductive success in bush crickets. Nature 307:361-363. Hamilton, W.D. 1964. The genetical evolution of social behavior I & II. Journal of Theoretical Biology 7:1-52. Huber. B.A. 2005. Sexual selection research on spiders: progress and biases. Biological Reviews 80:363-385. Jackson, R.R. 1980. The mating system of Phidippus johnsoni (Araneae, Salticidae): II. Sperm competition and the function of copulation. Journal of Arachnology 8:217-240. Johannesen, J., A. Hennig, B. Dommermuth & J.M. Schmeider. 2002. Mitochondrial DNA distributions indicate colony propagation by single matri-Iineages in the social spider Stegodyphus dumi- cula (Eresidae). Biological Journal of the Linnean Society 76:591-600. Jones, T.C. & P.G. Parker. 2000. Costs and benefits of foraging associated with delayed dispersal in the spider Anelosimus studiosus (Araneae: Theridiidae). Journal of Arachnology 28:61-69. Jones, T.C. & P.G. Parker. 2002. Delayed juvenile dispersal benefits both mother and offspring in the cooperative spider Anelosimus studiosus (Araneae: Theridiidae). Behavioral Ecology 13:142-148. Jones, T.C., S.E. Riechert, S.E. Dalrymple & P.G. Parker. 2007. Fostering model explains variation in levels of sociality in a spider system. Animal Behaviour 73:195-204. Klein, B.A., T.C. Bukowski & L. Aviles. 2005. Male residency and mating behavior in a subsocial spider. Journal of Arachnology 33:703-710. Lewis, W.M. Jr. 1987. The cost of sex. Pp. 33-57. In The Evolution of Sex and Its Consequences. (S.C. Steams, ed.). Birkhuser-Verlag, Basel. Matsumoto, T. 1993. The effect of the copulatory plug in the funnel- web spider Agelena limbata (Araneae: Agelenidae). Journal of Arachnology 21:55-59. Miller, J.A. 2007. Repeated evolution of male sacrifice behavior in spiders correlated with genital mutilation. Evolution 61:1301-1315. Miyashita, T. & H. Hayashi. 1996. Volatile chemical cue elicits mating behavior of cohabiting males of Nephila clavata (Araneae, Tetragnathidae). Journal of Arachnology 24:9-15. 532 THE JOURNAL OF ARACHNOLOGY Parker, G.A. 1984. Sperm competition and the evolution of animal mating strategies. Pp. 1-60. In Sperm Competition and the Evolution of Animal Mating Systems. (R.L. Smith, ed.). Academic Press, London. Parker, P.G., A. A. Snow, M.D. Schug, G.C. Booten & P.A. Fuerst. 1998. What molecules can tell us about populations: choosing and using a molecular marker. Ecology 79:361-382. Parker, P.G. & T.A. Waite. 1997. Mating systems, effective population size and conservation of natural populations. Pp. 243-261. In Behavioral Approaches to Conservation in the Wild. (J. Clemmons & R. Buckholz, eds.). Cambridge University Press, Cambridge, UK. Rayor, L.S. & L.A. Taylor. 2006. Social behavior in anrblypygids, and reassessment of arachnid social patterns. Journal of Arachnol- ogy 34:399-42 1 . Riechert, S.E. & R.M. Roeloffs. 1993. Evidence for and consequences of inbreeding in the cooperative spiders. Pp. 283-303. In The Natural History of Inbreeding and Outbreeding. (N.W. Thornhill, ed.). University of Chicago Press, Chicago. Rovner, J.S. 1996. Conspecific interactions in the lycosid spider Rabidosa rabida : the roles of different senses. Journal of Arachnology 24:16-23. Searcy, L.E., A.L. Rypstra & M.H. Persons. 1999. Airborne chemical communication in the wolf spider Pardosa milvina. Journal of Chemical Ecology 25:2527-2533. Schafer, M.A. & G. Uhl. 2002. Determinants of paternity success in the spider Pholcus phalangioides (Pholcidae: Araneae): the role of males and female mating behaviour. Behavioral Ecology and Sociobiology 51:368-377. Schneider, J.M., L. Fromhage & G. Uhl. 2005. Copulation patterns in the golden orb-web spider spider Nephila madagascariensis. Journal of Ethology 23:51-55. Shultz, Z. & S. Toft. 1993. Identification of a sex pheromone from a spider. Science 260:1635-1637. Smith, D.R. & R.H. Hagen. 1996. Population structure and interdemic selection in the cooperative spider Anelosimus eximius. Journal of Evolutionary Biology 9:589-608. Snow, L.S.E., A. Abdel-Mesih & M.C.B. Andrade. 2006. Broken copulatory organs are low-cost adaptations to sperm competition in redback spiders. Ethology 112:379-389. Snow, L.S.E. & M.C.B. Andrade. 2005. Multiple sperm storage organs facilitate female control of paternity. Proceedings of the Royal Society B-Biological Sciences 272:1 139-1 144. Trabalon, M., A.G. Bagneres & C. Roland. 1997. Contact sex signals in two sympatric spider species, Tegenaria domestica and Tegenaria pagana. Journal of Chemical Ecology 23:747-758. Trabalon, M„ G. Pourie & N. Hartmann. 1998. Relationships among cannibalism, contact signals, ovarian development and ecdysteroid levels in Tegenaria atrica (Araneae, Agelenidae). Insect Biochem- istry and Molecular Biology 28:751-758. Trivers, R.L. 1972. Parental investment and sexual selection. Pp. 136-179. In Sexual Selection and the Descent of Man, 1871- 1971. (B. Campbell, ed.). Aldine-Atherton Publishing Company, Chicago. Watson, P.J. 1986. Transmission of female sex pheromone by males in the spider Linyphia litigiosa Keyserling (Linyphiidae). Science 233:219-221. Watson, P.J. 1993. Foraging advantage of polyandry for female sierra dome spiders (Linyphia litigiosa , Linyphiidae) and assessment of alternative direct benefit hypotheses. American Naturalist 141:440-465. Watson, P.J. 1998. Multi-mating and female choice increase offspring size and growth in the spider Neriene litigiosa (Linyphiidae). Animal Behaviour 55:387-403. Williams, J.G.K., A.R. Kubelik, K.J. Livak, J.A. Rafalski & S.V. Tingey. 1990. DNA polymorphisms amplified by arbitrary primers are useful as genetic-markers. Nucleic Acids Research 18:6531-6535. Wilson. E.O. 1971. The Insect Societies. Belknap Press of Harvard University Press, Cambridge, Massachusetts. 548 pp. Yoward. P. & G. Oxford. 1996. Single palp usage during copulation in spiders. Newsletter of the British Arachnological Society 77:8-9. Manuscript received 17 December 2006, revised 6 March 2008. 2008. The Journal of Arachnology 36:533-537 Frequency and consequences of damage to male copulatory organs in a widow spider Michal Segoli: Department of Life Sciences, Ben-Gurion University of the Negev, Beer-Sheva, Israel. E-mail: msegoli@bgu.ac.il Yael Lubin: Mitrani Department of Desert Ecology, Jacob Blaustein Institute for Desert Research, Ben-Gurion University of the Negev, Sede Boqer Campus, Israel Ally R. Harari: Department of Entomology, Agricultural Research Organization, The Volcani Center, Bet Dagan, Israel, and Department of Life Sciences, Ben-Gurion University of the Negev, Beer-Sheva, Israel Abstract. Copulatory organ breakage, in which a portion of the male’s genitalia breaks off and remains in or attached to the female’s genitalia may represent a male strategy of high investment in a single mating. Such a strategy is expected when mating opportunities for males are limited and competition for females is high. We studied costs and benefits for males as a consequence of male organ breakage in the white widow spider (Latrodectus pallidus O. Pickard-Cambridge 1872). In order to estimate the frequency and consequences of such damage we provided each male with four virgin females simultaneously in an outdoors enclosure. We recorded male mating success and loss of the tip of the embolus (the male intromittent organ) inside the female’s genitalia for each male. In order to test the effect of the broken tip as a mating plug, we collected females from natural populations and observed the location of embolus tips inside their genitalia. We found that damage to the male organ was frequent but did not necessarily result in male sterility. From the field data, we found that the likelihood of a second embolus tip entering the spermatheca is significantly lower than that of the first tip, suggesting the possibility that the tip functions as a partial mating plug. Keywords: Male mating strategy, embolus tip, Latrodectus pallidus The reproductive success of a male is usually a function of the number of females he inseminates, especially when males produce numerous gametes and when little time and energy is spent on care of offspring. Under these circumstances each male is expected to mate with many females (Darwin 1871; Bateman 1948; Trivers 1972; Andersson 1994). Nevertheless, in some cases males invest highly in a single female. This investment may increase the male’s reproductive success in the current mating but dramatically reduce the number of matings that he can potentially achieve. Such a strategy can be promoted by evolutionary processes if the probability of encountering and mating with an additional receptive female is sufficiently low (Parker 1979; Buskirk et al. 1984; Elgar 1992; Simmons et al. 1992; Parker 1998; Andrade 2003) and if males strongly compete for females (Thornhill 1980; Fromhage et al. 2005). Copulatory organ breakage, in which a portion of the male’s genitalia breaks off during copulation is relatively common in spiders (Wiehle 1967; Breene & Sweet 1985; Foelix 1996; Schneider et al. 2001; Miller 2007). Broken organs inside the female’s genitalia may function as a mating plug to prevent fertilization by later arriving males, but it may also reduce the probability of the male fertilizing additional females (Foelix 1996). Thus, this trait may represent a male strategy of high investment in a single mating. To date, few studies have quantified the costs and benefits of male organ breakage. In the spider Nephila funestrata Thorell 1859, for example, males often damage both of their paired mating organs while copulating with a single virgin female (Fromhage & Schneider 2005). The occurrence of a male organ part inside the female’s genitalia was shown to reduce the number of copulatory insertions by a second male (Fromhage & Schneider 2006). Similarly, in Argiope bruennichi Scopoli 1772, males can use each copulatory organ once, and insertions into a previously used insemination duct were significantly shorter when the previous male had left parts of his genitalia inside the insemination duct (Nessler et al. 2007). In widow spiders ( Latrodectus ), the tip of one or both of the male’s intromittent organs (emboli) often breaks-off during copulation to be left inside the female’s genitalia (Levi 1959; Bhatnagar & Rempel 1962; Wiehle 1967; Kaston 1970; Berendonck & Greven 2002; Segoli et al. 2006). Males without embolus tips were assumed to be functionally sterile ( Bhatna- gar & Rempel 1962; Foelix 1996), but there is evidence that this is not always the case, as shown in L. mactans Fabricius 1775 (Breene & Sweet 1985) and L. hasselti Thorell 1870 (Snow et al. 2006). It was suggested that a tip inside the female’s spermatheca functions as a mating plug (Foelix 1996; Berendonck & Greven 2002); however, this was demonstrated only in L. hasselti. In this species first male sperm precedence was found when two males inseminated a single genital pore (Snow & Andrade 2005) and when the first tip was deposited in the entrance of the spermatheca (Snow et al. 2006). In several other Latrodectus species, however, more than one tip can be found inside the female’s spermathecae (Uhl 2002), suggesting that the embolus tip is not totally effective as a mating plug. In this study we investigated two aspects of the adaptive value and costs of damage to the male organ in the white widow spider, Latrodectus pallidus O. Pickard-Cambridge 1872: 1) future inseminating opportunities for males who have broken emboli (male sterility hypothesis) and 2) the risk of sperm competition (mating plug hypothesis). Males of this species suffer high extrinsic mortality and normally do not encounter more than one female in natural conditions, while females are often polyandrous (Segoli et al. 2006). Thus, males 533 534 THE JOURNAL OF ARACHNOLOGY that encounter a virgin female would benefit from blocking the spermatheca of their mate and thereby reducing or preventing access to future rivals. There is evidence that L. pallidus males invest highly in each mating: they cohabit in females’ webs longer than required for mating (Segoli et al. 2006), engage in an energetically demanding courtship display (M.S. personal observations), and are sometimes cannibalized by the female (Segoli et al. 2006). Thus, breakage of the male organ may be an integral part of the male mating strategy in this species even at the cost of a limited fertilization success in the future. In order to estimate the frequency and consequences of male organ breakage we asked the following questions: 1) How frequently do embolus tips break? 2) Does the loss of embolus tips prevent the male from remating? and 3) does the presence of a tip inside the female’s spermatheca reduce the probability of another tip entering the spermatheca? In order to answer the first two questions we conducted an experiment in which we provided 21 males each with four virgin females simultaneously in outdoor enclosures. Thus, each male had the opportunity to possibly mate with four females. For each male, we recorded fertilization success and the loss of embolus tips by recording the successful production of viable egg sacs in females and by recording the presence of the embolus tips in the genitalia of the females. In order to answer the third question, we collected females from natural populations and recorded the location of male tips inside their genitalia. METHODS The white widow spider (L. pallidus ) is common in the Negev desert of Israel (Levy 1998). We collected males and females from the Sede Teman area (31°17'N, 34°43'E) and Sayeret Shaked Park (31°16'N, 34°38'E) (northern Negev, Israel) in April 2003. Voucher specimens were deposited in the National Collection of Arachnids, Hebrew University of Jerusalem. Spiders were collected as juveniles or sub-adults and reared to maturity under lab conditions on the Sede Boqer Campus of Ben-Gurion University. Males were kept in plastic cups (200 cc) and fed weekly with Drosophila. Females were kept in terraria (10 X 20 X 15 cm) containing small dry shrubs, on which they constructed their webs. Females were fed weekly with flour beetles (larvae of Tenebrio molitor), crickets (Ache t a domestica), grasshoppers (Schistocerca sp.) and houseflies ( Musca domestica). Enclosure experiment. — We placed four adult virgin females with their webs in a square outdoor arena (135 cm length X 135 cm width X 50 cm height), one in each corner. Each arena was constructed from a wooden base and frame with plastic sheets as walls and a removable mesh cover. Females were not fed during the trial. Once the females repaired their webs (~1 day), we placed one adult naive male in the center of each arena (n — 21 replicates). The location and activity of the male (no movement, courting, in mating position) were recorded three times a day until it died. The number of daily observations was determined from preliminary observations, which indicated that males stay at least one day with each female. Females were measured and weighed at the end of each trial. They were kept until they produced seven egg sacs or until two months passed without laying eggs. Females that produced fertile egg sacs were assumed to have mated. This assumption is valid because mated females kept with adequate food rarely fail to produce fertile egg sacs (M.S. personal observations). The reproductive success of females was measured by the total number of eggs and by the number of fertile eggs (eggs that hatched) from the first five egg sacs. We used data from the first five egg sacs since most females lay 1- 3 egg sacs in the field and five egg sacs was the maximum observed in nature (M.S. personal observations). Post mortem we checked females’ spermathecae for the presence of embolus tips. Females have paired copulatory ducts, each leading to a spermatheca, and males have paired intromittent organs (emboli). During mating, the male inserts one embolus at a time into one of the female’s genital openings. Thus, a male may leave none, one, or both tips (one in each side) in the genitalia of a female. We obtained complete data on male embolus tips in the genital tracts of all four females from 15 trials. Spermathecae were examined by placing them in a 5% KOI I solution; after a week the tissue became transparent and the embolus tips were visible under a dissecting microscope (Berendonck & Greven 2002). For each trial we determined the order and number of females the male visited, which of the females he inseminated, the reproductive success of each female and finally, which females possessed embolus tips inside the spermathecae. Females collected from the field. — We collected 216 adult females from their webs at three locations in Israel: Goral Hills, near Lehavim (31°22'N, 34°49'E, n = 192), Kfar Edomim (31°49'N, 35°19'E, n = 13), and Sayeret Shaked (31°16'N, 34°38'E, n = 11) from March 1998 till September 1999. Egg sacs, if present, were left unharmed in the web. We dissected the females, removed their spermathecae and copulatory ducts, placed them in a 5% KOH solution and examined them for the presence of embolus tips as above. Since the number of tips in the right and left genitalia were correlated (Spearman rank correlation, n = 216, Rs = 0.632, P < 0.01), we considered only the right spermatheca and copulatory duct of each female, thereby avoiding pseudorep- lication. We compared cases in which one male tip was found in the female genitalia (genital duct + spermatheca) to cases in which two tips were found. We estimated the probability of a first embolus tip to enter the spermatheca as the percentage of females with an embolus tip located inside the spermatheca out of the total number of females with one embolus tip found in their right genitalia. We compared this with the probability of a second tip entering the spermatheca: the percentage of females with two tips inside the spermatheca, out of the total number of females with two tips in their right genitalia. We expected that if the first tip prevents the second tip from entering the spermatheca, the probability of finding a second tip inside the spermatheca would be lower than for the first tip. RESULTS Enclosure experiment, — After placement in the arena with the four females, males started courting one of the females. Courtship included the following behaviors: adding silk to the female’s threads, vibrating the web and cutting sections from the web. Mating was difficult to observe since it took place inside the female’s retreat. On the following days males were observed courting or standing motionless on the web, either inside or outside the retreat. The median time from the SEGOLI ET AL.— DAMAGE TO MALE CQPULATORY ORGAN 535 Number of mated females Figure i. — Number of males that mated with 0-3 females (n = 19 trials). Each male was provided with four virgin females simulta- neously in an outside enclosure. introduction of the male into the arena with females until the death of the male was 5 days (range 1-23 days, n = 21). One male escaped from his arena and entered another arena. He was returned to his arena after visiting one female in the adjacent arena. We excluded the two males from these two arenas from analyses of the number of mated females. Three males out of 19 did not inseminate any female, most of the males inseminated one or two females and two males inseminated three females (Fig. 1). The proportion of insem- inated females was higher among females that were visited first (77%) than among females visited later (41%) (Fisher’s exact test, n = 22 for first females and n = 29 for females visited later, P — 0.02). Three mated females died during the experiment and the remaining mated females produced seven egg sacs before the end of the experiment. The mean number of eggs per egg sac was 130 ± 30 (± SD, n = 30 females; averages of eggs per sac for each female were averaged over all females) and the number of hatched eggs was 103 ± 40. The total number of eggs that were produced by mated females in the first five sacs was not influenced by the number of embolus tips inside their spermathecae nor by mass, size, or age of females (GLM stepwise backward model, n = 25, P > 0.1 for all). The results were still not significant when considering hatched eggs only. Thus, there were no differences among the females in their reproductive success. Six out of 21 males (29%) were cannibalized by females. Cannibalism was observed directly or could be inferred from the transparent body of the dead male found on the female’s web. In five out of six cases the cannibalistic female did not produce egg sacs, indicating that cannibalism occurred before copulation, or that the female did not use the male’s sperm for fertilization. Data on the presence of male embolus tips in the genitalia of females and fertilization success are presented in Table 1. Three males did not lose any embolus tip with the first female they visited; nevertheless, one of these fertilized the female. Four males lost one embolus tip in the first mating. Two of these fertilized the first female only and the other two fertilized one and two additional females. Eight males lost both embolus Table 1. — Embolus tips inside spermathecae and fertilization of 1st, 2nd, and 3rd females visited by 15 males in the arena experiment. Numbers in columns represent the number of tips found in the female spermathecae. Shaded cells indicate that the female produced fertile egg sacs. # males First Second Third 2 0 0 0 1 0 0 0 2 1 0 0 4 2 0 0 3 2 0 0 1 2 0 0 1 1 1 0 1 1 1 0 tips in the first mating. Four of these fertilized the first female only, three fertilized an additional female, and one fertilized two additional females. Females collected from the field. — Forty-eight out of 216 females contained no embolus tip inside their right genitalia. In 86 cases out of the 95 females that contained one tip in their genitalia, the tip was placed inside the spermatheca, and in nine cases the tip was located in the genital duct. Thus, we estimated the probability for the first tip in the female’s genitalia to enter the spermatheca as 0.9. We found 58 females with two embolus tips inside their right genitalia. Out of these, in three cases both tips were located in the genital duct, in 28 the two tips were placed inside the spermatheca, and in 27 one tip was placed inside the spermatheca and the second was In the genital duct. Thus, we estimated the probability of a second tip to enter the spermatheca as 0.5. The likelihood of a first tip to enter the spermatheca was significantly greater than the likelihood of a second tip to enter the spermatheca (Fig. 2, Fisher’s exact test, P < 0.0001). Additionally, fourteen females contained three tips in their right genitalia. Of these, one female had no tips inside the right spermatheca, 8 had one tip, two had two tips and three had three tips. Finally, one female contained five embolus tips, four of which were found inside the spermatheca. Total # of tips in right genitalia (spermatheca + dye!) Figure 2. — Embolus tips found in the spermatheca alone and in the entire genitalia: percentage of females with no embolus tips (white section), one embolus tip (gray section) or two embolus tips (black section) inside their right spermatheca, out of field-collected females with either one (n = 95) or two tips (n = 58) in their right genitalia (spermatheca + genital duct). 536 THE JOURNAL OF ARACHNOLOGY DISCUSSION In this study we estimated the frequency and consequences of damage to the male copulatory organ in the white widow spider L. pallidus. We found that damage to the male organ was frequent but did not necessarily result in male sterility. We showed that the occurrence of a male’s embolus tip inside the female’s spermatheca functions as a partial mating plug: it probably obstructs but does not always prevent the entrance of an additional tip into the spermatheca. In contrast to our study, male sterility following damage to the male copulatory organ has been demonstrated in several spider species. For example, in Argiope keyserlingi Karsch 1878, experimental removal of one copulatory organ prevent- ed males from copulating with more than one female, suggesting that males can use each of their paired organs only once (Herberstein et al. 2005). In Nephila funestrata, 95% of the males mating with a virgin female had a damaged organ that probably prevented them from remating (Fromhage & Schneider 2005). In widow spiders the loss of an embolus tip inside the female genital tract was previously assumed to result in functional sterility of the male (Bhatnagar & Rempel 1962). Breene & Sweet (1985), however, found that some males of a congener (L. mactans) were able to successfully inseminate three females suggesting that males either do not always lose their tips or that they can inseminate in spite of embolus breakage. In L. hasselti , males are normally sterile after mating (Andrade & Bant a 2002), but when tips were cut experimentally males were able to inseminate additional females (Snow et al. 2006). Thus, the loss of embolus tips alone cannot be responsible for the post-mating sterility in L. hasselti. In our study we found that at least one male mated and inseminated a female without losing any embolus tip and five males inseminated one or two females after losing both tips in previous matings. Thus, we suggest that the loss of embolus tips in L. pallidus is common, but does not prevent the male from fertilizing additional females. Although damage to the copulatory organ in L. pallidus was not an absolute constraint on the male’s reproductive success, only a few males (2 out of 19) inseminated more than two females. This suggests that insemination with a broken embolus is mechanically difficult and is less likely to be successful than insemination with an intact embolus. Addi- tionally, in the absence of tips, males may have difficulties filling their emboli with sperm (sperm induction) and therefore low fertilization success may result from sperm depletion rather than an inability to transfer sperm (Snow et al. 2006). However, it is not yet known whether white widow males refill their emboli between mating attempts. Finally, insemination with a broken embolus may be especially difficult when mating with an already mated female with a plugged spermatheca. If so, embolus breakage may still carry a cost for males in mating systems where sperm competition exists. Although males were not competing for females in this experiment, there is evidence that embolus breakage may give the males an advantage in sperm competition. Most of the males that lost both tips (8 out of 10) left them in each of the two spermathecae of the first female that they mated, indicating that they had mated with her twice. However, there was no difference in the reproductive success of females with one or two tips in their spermathecae. A similar result was obtained in a study of L. hasselti where repeated mating did not increase the probability of successful fertilization nor the number of offspring produced in successful matings (Andrade & Santa 2002). We suggest that males leave both tips in order to protect both of the female’s spermathecae from future insemination by rival males. The analysis of spermathecae from females collected in the field further supports the view that the broken embolus functions as a partial mating plug. The probability of a first embolus tip entering the spermathecae was significantly higher (90%) than that of the second tip (50%). It is also possible that a second tip replaced the first, but this is unlikely considering the narrow entrance to the spermatheca (Beren- donck & Greven 2002). However, the results also suggest that the tip is not totally effective as a plug: in half of cases a second tip did enter the spermatheca, and in four cases more than two tips entered the spermatheca. In contrast, in a study of L. hasselti it was shown that in —90% of the cases where two males inseminated the same genital pore, the second tip did not enter the spermatheca resulting in a first male sperm priority (Snow et al. 2006). Although it is difficult to compare the results of this controlled experiment with our field data, it implies that the plug in L. hasselti is more efficient than in L. pallidus. From an evolutionary point of view, the differences in the efficiency of the plug between species may reflect an arm-race between males and females over control of paternity. In this light it would be interesting to compare the efficiency of the plug in different Latrodectus species in relation to mating opportunities, effective sex ratio, and sexual canni- balism. In contrast to embolus tip breakage, sexual cannibalism does not seem to be an integral part of the male mating strategy in L. pallidus. In L. hasselti , males initiate cannibalism by placing their abdomen in front of the female’s mouthparts during copulation: cannibalized males copulate longer and cannibalistic females are less likely to remate (Andrade 1996). Similar sacrificial behavior was also observed in L. geome- tricus C.L. Koch 1841 (Segoli et al. 2008). In L. pallidus , however, most of the cannibalized males (5 out of 6) did not fertilize the cannibalistic female and thus could not benefit from cannibalism. This illustrates the distinction between male sacrifice behavior as an adaptive strategy and cannibalism as an unavoidable consequence of mating with a dangerous partner. In conclusion, damage to male copulatory organs is consistent with a male strategy of high investment in a single female. Embolus damage does not necessarily result in male sterility and may provide some paternity advantage over subsequent males. This benefit will be expressed only when females mate multiply and when mating opportunities are limited for males, as is the case in the mating system of white widow spiders (Segoli et al. 2006). ACKNOWLEDGMENTS We thank Iris Musli, Efrat Gavish, Alexei Maklakov, Dinesh Rao, Yael Teleman, Moran Segoli, Tamar Keasar, and Jutta Schneider for discussions and assistance, and two anonymous referees for valuable comments on the manuscript. This is publication no. 613 of the Mitrani Department of Desert Ecology. SEGOLI ET AL.— DAMAGE TO MALE COPULATORY ORGAN 537 LITERATURE CITED Andersson, M. 1994. Sexual Selection. Princeton University Press, Princeton, New Jersey. 559 pp. Andrade, M.C.B. 1996. Sexual selection for male sacrifice in the Australian redback spider. Science 271:70-72. Andrade, M.C.B. 2003. Risky mate search and male self-sacrifice in redback spiders. Behavioral Ecology 14:531-538. Andrade, M.C.B. & E.M. Banta. 2002. Value of male remating and functional sterility in redback spiders. Animal Behaviour 63:857— 870. Bateman, A.J. 1948. Intra-sexual selection in Drosophila. Heredity 2:349-368. Berendonck, B. & H. Greven. 2002. Morphology of female and male genitalia in Latrodectus revivensis with regard to sperm priority patterns. Pp. 157-167. In European Arachnology 2000. (S. Toft & N. Scharff, eds.). Aarhus University Press, Aarhus, Denmark. 358 pp. Bhatnagar, R.D.S. & J.G. Rempel. 1962. The structure, function, and postembryonic development of the male and female copulatory organs of the black widow spider Latrodectus curacaviensis (Muller). Canadian Journal of Zoology 40:465-510. Breene, R.G. & M.H. Sweet. 1985. Evidence of Insemination of multiple females by the male black-widow spider, Latrodectus mactans (Araneae, Theridiidae). Journal of Arachnology 13:331— 335. Buskirk, R.E., C. Frohlich & K.G. Ross. 1984. The natural-selection of sexual cannibalism. American Naturalist 123:612-625. Darwin, C.R. 1871. The Descent of Man and Selection in Relation to Sex. The Modern Library, Random House, New York. 743 pp. Elgar, M.A. 1992. Sexual cannibalism in spiders and other invertebrates. Pp. 128-155. hi Cannibalism: Ecology and Evolu- tion Among Diverse Taxa. (M. Elgar & B. Crespi, eds.). Oxford University Press, Oxford, UK. Foelix, R. 1996. Biology of Spiders. Second edition. Oxford University Press, New York. 330 pp. Forster, L.M. 1992. The stereotyped behavior of sexual cannibalism in Latrodectus hasselti Thorell (Araneae, Theridiidae), the Austra- lian redback spider. Australian Journal of Zoology 40:1-11. Fromhage, L., M.A. Elgar & J.M. Schneider. 2005. Faithful without care: the evolution of monogyny. Evolution 59:1400-1405. Fromhage, L. & J.M. Schneider. 2005. Safer sex with feeding females: sexual conflict in a cannibalistic spider. Behavioral Ecology 16: 377-382. Fromhage, L. & J.M. Schneider. 2006. Emasculation to plug up females: the significance of pedipalp damage in Nephila fenestrata. Behavioral Ecology 17:353-357. Herberstein, M.E., K.L. Barry, M.A. Turoczy, E. Wills, C. Youssef & M.A. Elgar. 2005. Post-copulation mate guarding in the sexually cannibalistic St Andrew’s Cross spider (Araneae Araneidae). Ethology Ecology & Evolution 17:17-26. Kaston, B.J. 1970. Comparative biology of American black widow spiders. Transcriptions of the San Diego Society of Natural History 16:33-82. Levi, H.W. 1959. The spider genus Latrodectus (Araneae, Theridi- idae). Transactions of the American Microscopical Soceity 78: 7-43. Levy, G. 1998. Fauna Palestina Arachnida III: Araneae: Theridiidae. Academy of Sciences and Humanities, Jerusalem, Israel. 264 pp. Miller, J.A. 2007. Repeated evolution of male sacrifice behavior in spiders correlated with genital mutilation. Evolution 61:1301-1315. Nessler, S.H., G. Uhl & J.M. Schneider. 2007. Genital damage in the orb-web spider Argiope bruennichi (Araneae: Araneidae) increases paternity success. Behavioral Ecology 18:174-181. Parker, G.A. 1979. Sexual Selection and Sexual Conflict. Pp. 123-166. In Sexual Selection and Reproductive Competition in Insects. (M.S. Blum & N.A. Blum, eds.). Academic Press, New York. Parker, G.A. 1998. Sperm competition and the evolution of ejaculates: towards a theory base. Pp. 3-54. In Sperm Competition and Sexual Selection. (T.R. Birkhead & A.P. Moller, eds.). Academic Press, London. Schneider, J.M., M.L. Thomas & M.A. Elgar. 2001. Ectomised conductors in the golden orb-web spider, Nephila plumipes (Araneoidea): a male adaptation to sexual conflict? Behavioral Ecology and Sociobiology 49:410-415. Segoli, M., R. Arieli, P. Sierwald, A.R. Harari & Y. Lubin. 2008. Sexual cannibalism in the brown widow spider (Latrodectus geometricus). Ethology 1 14:279-286. Segoli, M„ A.R. Harari & Y. Lubin. 2006. Limited mating opportunities and male monogamy: a field study of white widow spiders, Latrodectus pallidus (Theridiidae). Animal Behaviour 72: 635-642. Simmons, L.W., R.J. Teale, M. Maier, R.J. Standish, W.J. Bailey & P.C. Withers. 1992. Some costs of reproduction for male bush- crickets, Requena verticals (Orthoptera, Tettigoniidae): allocating resources to mate attraction and nuptial feeding. Behavioral Ecology and Sociobiology 31:57-62. Snow, L.S.E., A. Abdel-Mesih & M.C.B. Andrade. 2006. Broken copulatory organs are low-cost adaptations to sperm competition in redback spiders. Ethology 112:379-389. Snow, L.S.E. & M.C.B. Andrade. 2005. Multiple sperm storage organs facilitate female control of paternity. Proceedings of the Royal Society B-Biological Sciences 272:1139-1144. Thornhill, R. 1980. Sexual selection within mating swarms of the lovebug, Plecia nearctica (Diptera, Bibionidae). Animal Behaviour 28:405-412. Trivers, R.L. 1972. Parental investment and sexual selection. Pp. 136-179. In Sexual Selection and the Descent of Man, 1871- 1971. (B.G. Campbell, ed.). Aldine-Atherson, Chicago. Uhl, G. 2002. Female genital morphology and sperm priority patterns in spiders (Araneae). Pp. 145-156. In European Arachnology 2000. (S. Toft & N. Scharff, eds.). Aarhus University Press, Aarhus, Denmark. Wiehle, H. 1967. Steckengebliebene Emboli in den Vulven von Spinnen (Aracfa., Araneae). Senckenbergiana Biologica 48:197- 202. Manuscript received 13 May 2007, revised 30 April 2008. 2008. The Journal of Arachnology 36:538-544 Molting interferes with web decorating behavior in Avgiope keyserlingi (Araneae, Araneidae) Andre Walter1, Mark A. Elgar2, P. Bliss1 and Robin F. A. Moritz1: 'Institut fur Biologie, Martin-Luther-Universitat Halle-Wittenberg, Hoher Weg 4, D-06099 Halle (Saale), Germany. E-mail: andre.walter@zoologie.uni-halle.de; department of Zoology, University of Melbourne, Melbourne, Victoria 3010, Australia Abstract. Various orb weaving spiders decorate their webs with extra silk structures. In the araneid genus Avgiope, these web decorations consist of flimsy aciniform silk threads arranged in zig zag shaped bands. The adaptive value of these structures is still unclear and controversy over a suite of possible functional explanations persists: the high variation of web decoration adds further uncertainty. Web decorations can differ in shape, size, and frequency across species and even within species. Physiological processes may influence individual variation in web decorating behavior. Molting events are major physiological transitions combined with fundamental alterations of the metabolic state of the spiders. For gaining new insights into possible proximate mechanisms driving web decorating behavior, we observed subadult Avgiope keyserlingi Karsch 1878 females in the laboratory and registered the individual variation of web decorations associated with the maturity molt under laboratory conditions. We found substantial individual variation of web decorations of A. keyserlingi. The most striking result was that subadult spiders built dramatically oversized decorations prior to the last molt. Since aciniform silk is used for both constructing web decorations and immobilizing prey we suggest that these extensive decorations might provide a store for the swift replenishment of aciniform silk after the molt. High silk recycling rates make temporary outsourcing less costly and facilitate a rapid resumption of prey capture following lost foraging opportunities during the molting phase. Thus, we argue that the solution of the riddle of web decorations might reside in the physiology of molting spiders. Keywords: Orb-web spiders, web decorations, maturity molt, gland regulation Web decorating is a characteristic behavior of various orb weaving spiders (Robinson & Robinson 1973; Edmunds 1986; Bruce 2006), yet the possible functional explanations remain controversial despite extensive investigations (Herberstein et al. 2000; Eberhard 2003; Bruce 2006). “Web decorations” (first mentioned as such by McCook 1889, but also called “stabilimenta” by Simon 1895 and many modern authors) in the araneid genus Avgiope consist of numerous flimsy acini- form silk threads (Peters 1993), mostly arranged in zig zag shaped bands (Bruce 2006). Although web decorations of Avgiope are considered as prey attractants by some (Craig & Bernard 1990; Tso 1996; Bruce et al. 2001; Li 2005; Cheng & Tso 2007), this view is not unanimous and alternative functional explanations include anti-predator devices (Ewer 1972; Schoener & Spiller 1992; Blackledge & Wenzel 2001); advertisement for web protection (Eisner & Nowicki 1983; Kerr 1993; Blackledge & Wenzel 1999); thermoregulation (Humphreys 1992); mechanical support [Robinson & Robin- son 1970; see also Watanabe 2000 for Octonoba sybotides (Bosenberg & Strand, 1906), Uloboridae]; and acting as a molting platform (Robinson & Robinson 1973, 1978). In the rapidly growing literature on this topic, tests for non-visual functions are clearly underrepresented (Bruce 2006). In particular the potential relationship between physiological processes and web decorating behavior has been addressed in only a very few studies (e.g., Peters 1993; Tso 2004; Walter et al. 2008a). Typically, decorating behavior in species of Avgiope is highly variable (Bruce & Herberstein 2005) and web decora- tions can differ in shape (number and arrangement of zig zag bands), size and frequency (Lubin 1975; Edmunds 1986; Nentwig & Heimer 1987; Schoener & Spiller 1992). One problem for determining the adaptive value of web decora- tions stems from this high variation (Robinson & Robinson 1974), which occurs across species and within species at both the population and individual level (Herberstein et al. 2000; Starks 2002; Bruce & Herberstein 2005; Rao et al. 2007). Most studies explore the adaptive significance of these structures (e.g., Blackledge 1998; Craig et al. 2001), although phyloge- netic analyses of web decoration patterns suggests that interspecific variance shows weak homologies at best and yields phylogenetically feeble signals (Herberstein et al. 2000; see also Scharff & Coddington 1997). We agree with Eberhard (2003) that an accumulation of single “experiments per se ... are no guarantee of reliable conclusions.” Thus, understand- ing the intra-individual variance of web decorations in detail is necessary before embarking on the interpretation of web decorating behavior in general. The production of web decoration is governed by an enhanced activity of the silk glands and hence physiological processes are expected to impact web decorating behavior (Tso 2004; Walter et al. 2008a). The major physiological transitions in the life history of spiders are the repeated molting events. Molting requires a drastic change of anabolic and metabolic biochemical pathways requiring fundamental alterations of the physiological state of the animal. Apart from hormonal changes (Bonaric 1987; Foelix 1996; Craig 2003), molts are particularly vulnerable events in the life of spiders, in terms of both increased physiological stress (Pulz 1987; Vollrath 1987a) and increased risk of predation (Tolbert 1975; Tanaka 1984; Vollrath 1987b; Baba & Miyashita 2006). It would, therefore, be surprising if web decoration behavior was not affected by molting. Indeed several studies suggest that molting might have profound effects on the web decorating activity of Avgiope (Robinson & Robinson 1970, 1973; Edmunds 1986; Nentwig & Heimer 1987). Yet if we observe consistent changes in the patterns of decoration behavior associated with the molting process, this might 538 WALTER ET AL.— WEB DECORATING BEHAVIOR IN ARGIOPE KEYSERLINGI 539 provide insights into the proximate mechanisms driving web decoration and their potential adaptive value. METHODS Study species and experimental design. — We chose the St. Andrew’s cross spider, Argiope keyserlingi Karsch 1878, to study the variation in web decoration under highly controlled laboratory conditions. This orb web spider is distributed along the east coast of Australia (northern Queensland to Victoria in the south), building their webs between branches and leaves of bushes, e.g., in parks and gardens. Argiope keyserlingi is a well studied species in terms of its natural history (Rao et al. 2007), its sexual cannibalism (Elgar et al. 2000; Herberstein et al. 2005) and its web decorating behavior (Herberstein 2000; Bruce et al. 2001, 2005; Herberstein & Fleisch 2003). St. Andrew’s cross spiders typically build cruciate web decorations consisting of up to four zig zag bands forming a large “X” in the orb web (Rao et al. 2007). This allows an unambiguous interpretation of deviations from the “complete cross.” We collected 55 subadult female spiders in Ku-ring-gai Bicentennial Park (West Pymble/Sydney, Australia) and trans- ferred them individually to Perspex frames (58 cm X 58 cm X 15 cm) in the laboratory, where they were kept under natural light conditions. Every other day, each spider was fed with one blowfly ( Lucilia spp.). At this life stage the spiders are still of a similar size as the blowflies; thus, it has turned out in preliminary observations that this feeding regime is sufficient to keep spiders “well-fed.” At this same time, each web was moistened with five shots from a water spray. Given that spiders typically build a new web each day, we recorded daily the number of decoration bands (shape) and decoration size to assess the variation of web decorating behavior within a total observation period of 30 days. We estimated the size of the web decoration by computing a trapezium area similar to Tso (1999): (a+c)/2 X h (a and c = upper and lower width of zig zag bands, h = height of zig zag bands, see Fig. 1 ). Additionally, we quantified the size of all newly built webs following Herberstein & Tso (2000): (dv/2) X (dh/2) X n (dv = vertical and dh = horizontal diameter of the capture area, see Fig. 1) and measured the spider body size (length from clypeus to the end of the opisthosoma). Voucher specimens were deposited in the Entomological Collection of the Martin-Luther-University Halle-Wittenberg (Zoological Insti- tute), Germany (identification number 2569). Statistical analyses. — We used STATISTICA® (version 6.0) for all statistical analyses including the paired /-test to evaluate differences in the sizes of decorated and undecorated webs. Chi square-tests and /-tests were used to detect differences in the proportion of decorated web parts and constructed decoration patterns. Web and web decoration sizes prior, during, and after molting events were analyzed with an ANOVA. Pearson-correlations were computed between web size and decoration size. RESULTS Web decorating frequency. — All females could be observed over the whole 30 day period. Forty-six of the 55 subadult A. keyserlingi molted to maturity within this time. The spiders constructed new webs every second day (mean 2.29 ± 0.07 SE day). Typically, the new web decorations were built together with new webs, and therefore the decorating activity mostly Figure 1 . — Web and decoration measurements from the webs of A. keyserlingi-. Left: determination of the size of capture areas (including hub region), dh = horizontal diameter, dv = vertical diameter; Right: determination of decoration band sizes, a = upper width of the band, c = lower width of the band, h = length of the band (lettering after trapezium formula). followed an equal rhythm (a mean value of every 2.37 ± 0.37 SE day). The few exceptions were all in the context of molting events (see below). However, 233 (37.5%) of all newly built webs ( n = 622) did not contain a web decoration. Many spiders occasionally failed to decorate their webs, but only Five animals (9.1%) never built a web decoration at all during the observation period. Web size. — The spiders more than doubled the catching area of their webs within the 30 day observation period. The mean size of the first web we measured was 635.30 ± 44.46 SE cm2 (n — 55) and the mean size of the last measured web was 1630.61 ± 21.99 SE cm2 (n = 55). Over the whole observation period, undecorated webs were significantly larger than decorated webs, ranging from 625.21 ± 56.88 SE cm- to 646.54 ± 70.57 SE cm2 at the beginning of to the experiment to between 1563.21 ± 42.56 SE cm2 and 1700.50 ± 37.42 SE cm2 at the end of the period (paired t-test: / = 2.1 1, P < 0.05). However, the mean decoration size did not significantly change over time, and ranged from 55.17 ± 72.72 SE mm" (n = 55) at the beginning to 46.25 ± 54.06 SE mm2 (n = 55) at the end of the observation period (Pearson, r — 0.04, P — 0.29). We found a significant positive correlation between spider size and web area (Pearson, r~ = 0.31, P < 0.01; n = 621). In contrast, we found no significant correlation between spider size and web decoration size. Consequently, the size of the decorated web area in relation to the total web decreased over time. Variation of web decorating behavior. — The variation in web decoration shapes was very high and the “typical” cruciate type was rarely constructed (Fig. 2); females of A. keyserlingi most often constructed single arm decorations (n = 47 spiders in 65.13%, n = 282 observations), and decorations with two (/? = 31 spiders in 24.48%, n — 106 observations), three (n — 10 spiders in 5.54%, n = 24 observations), or four (n = 15 spiders in 4.85%, n = 21 observations) zig zag bands were less frequent. In all partial cross shapes (one to three arms), the bands were added to the lower web half significantly more often (85.2% vs. 14.8%, n = 50; X2 = 9.12, P < 0.01). There 540 THE JOURNAL OF ARACHNOLOGY Number of web decoration bands (decorations, n=433) Figure 2. — The variation in web decoration patterns of A. keyserlingi females under laboratory conditions. Partial cross shapes (one to three decoration bands = number 1-3 in the diagram) are more frequent than the typical cruciate shape (number 4). was also strong intra-individual variance; most spiders (65.45%, n — 36) altered the web decoration pattern up to four times over the observation period: Thirteen spiders (23.64%) altered their web decoration pattern once; nine (16.36%) altered the pattern twice; six individuals (10.51%) altered it three times; and three spiders (5.46%) altered the decoration four times. Only 19 spiders (with 34.55% signifi- cantly less, X2 = 9.55, P < 0.01) constructed the same number of arms within the observation period and five individuals (9.09%) built no web decoration at all. These latter spiders also had a significantly lower web decorating frequency (new decoration every 3.7 ± 4 SE days) than individuals that constructed more variable shapes over time (new decoration every 1.5 ± 0.9 SE days, n = 31; t-test: t = 2.91, P < 0.01). Web decorating behavior in the context of molting events. Within the 30 day observation period 46 of 55 subadult spiders molted to maturity. Spiders suspended the two-day web building rhythm a few days before molting, and on average 3.3 ± 1.6 SE days elapsed between the ‘last” web building and the start of the molt. The mean interval between constructing the “last” web decoration prior to the final molt into sexual maturity (2.8 ± 1.5 SE days) was also longer than the mean decorating interval at other times (every 2.37 ± 0.37 SE day, see above). The molting events coincided with an increase in overall web size: the web size had increased by 19% (mean +260 cm2) in the ten day period after the molt (from 1080.37 ± 42.9 SE cm2, n = 101 prior to the molt to 1340.36 ± 25.56 SE cm2, n = 205; paired t-test: t = —4.88, P < 0.01). Ten spiders (22%) added a new web decoration to an old web prior to the molt. The change in web decorating and web building frequency was exclusively observed in combination with molting events, and the most conspicuous change was the dramatic increase in the web decoration size (Fig. 3) during the pre-molting phase (last subadult webs). The size of the “regular” web decorations, both in penultimate webs before and in the first webs after the molt, were significantly smaller (68.78 ± 10.45 SE mm2, n = 43 vs. 58.39 ± 9.24 SE mm2, n = 45) than those constructed directly in the last web before molting (21 1.74 ± 35.94 SE mm2, n = 46; ANOVA: F = 14.36, P < 0.01). The “supersized” decorations of the molting webs were characterized by a partial loss of the typical zig zag look (Fig. 3, right). Moreover, these peculiar decoration bands overlapped in the hub region of the web, which was never observed in intermolt webs. Finally, only one individual molted in a web without a web decoration. All in all, individuals of A. keyserlingi reduced their web building frequency (Fig. 4A) and increased the size of their web decorations prior to their final molt to sexual maturity (Fig. 4B). DISCUSSION Although individuals of A. keyserlingi usually build cruciate web decorations (Rao et al. 2007) consisting of up to four zig zag-shaped silk bands (Bruce 2006), we observed substantial individual variation in web construction and decorating behavior in A. keyserlingi in our study. Web size strongly correlated with the spider’s size and larger females built larger webs. Moreover, we could confirm previous reports by Hauber (1998) and Craig et al. (2001 ) on a negative correlation between web size and decoration size. Undecorated Argiope webs were larger than decorated ones. Since we kept the feeding regime constant, this might indicate a tradeoff between web size and decoration as suggested by Craig et al. (2001). Although web size was positively correlated with spider size, larger spiders did not build larger web decorations. Consequently, the relative decoration area of the web decreased over time, which may reflect previous reports of reduced web decorating behavior in later adult stages of Argiope spiders (Peters 1953; Edmunds 1986; Nentwig & Heimer 1987). The intra-individual variation in the shape of the decoration was remarkably high. Very few spiders consistently built only one particular pattern. An explanation for this may be given by the results of Craig et al. (2001) on Argiope argentata (Fabricius 1775). They argue that individ- ual decoration patterns have a genetic component and any variation represents the influence of ecological conditions. Most spiders in our study alternated the web decoration type, some individuals up to four times. Although this high variation may have been affected by the laboratory condi- tions it has also been observed in many other Argiope species (e.g., Blackledge 1998 in A. aurantia Lucas 1833 and A. trifasciata (Forskal 1775); Hauber 1998 in A. appensa (Walckenaer 1842); Seah & Li 2002 in A. versicolor (Doleschall 1859); Bruce & Herberstein 2005 in A. picta L. Koch 1871 and A. aetherea (Walckenaer 1842)). Argiope keyserlingi females in our study regularly rebuilt their orb webs every second day, and the web decorating frequency followed this rhythm. The only exceptions occurred on those days leading up to the commencement of the final molt to sexual maturity. During this time, some spiders added web decorations to their old webs. Typically, Argiope spiders do not rebuild orb webs several days before they molt to maturity (Robinson & Robinson 1978; Nentwig & Heimer WALTER ET AL.- WEB DECORATING BEHAVIOR IN ARGIOPE KEYSERLINGI 541 Figure 3. — The “regular” web decoration (left) and the “supersized” web decoration (right) of A. keyserlingi. 1987; Eberhard 1990). Robinson & Robinson (1973) suggest that a tradeoff between silk production and the biosynthetic efforts in preparation of the molt provides an adaptive explanation for this phenomenon. However, the frequency of web decorating prior to molting did not decline, despite the reduction in web building, because some spiders added new decorations to already existing, old, and dilapidated webs. Indeed, the dramatically oversized decorations that spiders built prior to the molt (Fig. 4) were the most conspicuous difference to the intermolt webs of subadult and the webs of adult individuals. The phenomenon that spiders build more frequent and/or more perfect web decorations prior to the molt was already observed in A. argentata and A. savignyi Levi 1968 in the laboratory (Nentwig & Heimer 1987). Moreover, Edmunds (1986) noticed larger and denser decorations prior to moltings in a wild population of A. Jlavipalpis (Lucas 1858). These anecdotal reports, however, have never been empirically quantified. In our study we could show that web decorations in A. keyserlingi were three times larger shortly before the maturation molt and did not correspond with the individual variation in decoration shape. Decoration size decreased to the intermolt level immediately after the molt. Consequently, very large decorations were thus directly linked to the molting procedure. Do our findings contribute to resolving the controversy over the adaptive significance of web decorations (see Bruce 2006)? Web decorations have been discussed in a variety of contexts, including in the context of prey attraction (Herber- stein 2000; Herberstein & Fleisch 2003; Li 2005). Although we cannot exclude (his explanation for decorations in regular webs, the observed increase in web decorating activity in A. keyserlingi prior to the molt is not predicted by this hypothesis. Spiders decrease their foraging efforts during the pre-molt phases (Higgins 1990), presumably because there is little opportunity to consume food during molting. Nevertheless, web decorations may provide particular me- chanical support for orb webs (Simon 1895) during the molting phase (Robinson & Robinson 1970, 1973, 1978; Nentwig & Heimer 1987). Higgins (1990) argued that the web decorations of Nephila clavipes (Linnaeus 1767) (Nephilidae) help prevent the spiders contacting the sticky spiral, which could interrupt the moiling procedure by hindering individ- uals from freeing themselves from the old exoskeleton. Since molting events are generally vulnerable phases in the life of a spider (Robinson & Robinson 1973; Baba & Miyashita 2006) the potentially protective properties of web decorations may be relevant in preserving the integrity of the web during the molt (Horton 1980; Eisner & Nowicki 1983; Kerr 1993). Additionally, the potential protection against predators (Eberhard 1990; Schoener & Spiller 1992; Blackledge & Wenzel 2001) would also predict an increase in decorating investment because spiders are especially vulnerable to predators during the molt or shortly afterwards (Tanaka 1984; Baba & Miyashita 2006). 542 THE JOURNAL OF ARACHNOLOGY O «D i- *“ II c r 0 3 |f Q. O .£ <5 > £ •S s it fl Web decoration size hk. t v hJ \ V * J A \ A -10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8 9 10 Day Web building frequency Figure 4. — A. The size of the web decoration of A. keyserlingi dramatically increases prior to the maturation molt and then returns to the level prior to the molting event; B. Web building frequency decreases prior to the last molt; dotted line: day of molting (= day 0). Shortly after a molt spiders are vulnerable to desiccation due to the slowly sclerotizing exoskeleton. In this phase it must be particularly important to balance the hygric status through water ingestions. In this context (large) web decorations might be practical tools: some Argiope spiders directly ingest water from parts of their web decorations (Olive 1980; Walter et al. 2008a). Since Argiope spiders can also successfully molt on webs without a web decoration (Nentwig 1986; own observations) the adaptive effects of the decorations may play a subsidiary role. Instead, the increase in web decoration investment may proximately derive from direct physiological processes, particularly since resource allocations directly influence interactions between molting, silk composition and web building behavior (Townley et al. 2006). Thus, it might be necessary to “outsource” a certain amount of nutrients for optimizing the molting procedure. Higgins & Rankin (2001) showed that “well-fed” individuals of the orb weaving spider N. clavipes more often suffer from molting failures when exceeding a critical pre-molt mass. They concluded that this might be the cost for the ability of rapid growth based on an almost non-limited food intake in this species. This may also be relevant for the rapid growth of Argiope spiders. Outsourcing body mass in the form of silk proteins may ensure an “optimal” molt-weight. In this context it is possible that N. clavipes builds web decorations only shortly before a molt (Higgins 1990). Conversely, a molt is always combined with a loss of body mass (through the failure to consume exuvia) (Hutchinson et al. 1997), and outsourcing silk proteins may allow spiders to minimize nutrient waste. The link between the increase of web decorating behavior and moltings might also be explained by a requirement to outsource specific, physiologically important compounds that would be otherwise metabolized during the molting proce- dure or the non-foraging days shortly before and after the molt. Such allocation occurs for different compounds in several spider species [e.g., choline, Higgins & Rankin 1999 for N. clavipes and Townley et al. 2006 for Argiope trifasciata and A. aurantia; GABamide, Townley & Tillinghast 1988 for Araneus cavaticus (Keyserling 1881)]. Perhaps the enlarged decoration simply provides a storage area for the silk proteins themselves. The aciniform decoration silk is also used for immobilizing prey (Peters 1993; Tso 2004). Thus, web decorating might be crucial for maintaining a certain level of activity in the aciniform glands for an optimal perfor- mance of Argiope' s typical “wrap attack” strategy of prey capture (Olive 1980; Tso 2004; Walter et al. 2008b). After molting, spiders must swiftly resume capturing prey to compensate for lost foraging opportunities of the previous days. For subsequent capture events, Argiope requires large amounts of wrapping silk that has to be newly synthesized after the molt. Since several types of silk glands are remodeled during a molt, they may not be fully operative in the days immediately after the molt (Townley et al. 2006). If this is also true for the glandulae aciniformes , the extensive web decorations may provide an ideal store of the crucial silk components, allowing the swift replenishment of the acini- form silk following molting. The highly efficient recycling of web parts (Peakall 1971) thereby clearly reduces the costs of silk production (Janetos 1982; Opell 1998) by reusing the relevant amino acids. To confirm the physiological background of our observa- tions, further studies should concentrate on the impact of different metabolic processes on the web decorating behavior prior to moltings, with a focus on those spiders that nonetheless molt without decorations. However, irrespective of the actual ultimate adaptive mechanisms of web decora- tions, it seems that these structures may play a more specific role in the molting web than in the regular capture web in Argiope. Given the large size of the molting decorations in contrast to relatively small and highly variable decorations in regular webs, it may well be that the clue to solving the riddle of these structures lies in the physiology of the molting spider. ACKNOWLEDGMENTS We thank two anonymous reviewers for their useful comments. This work is supported, in part, by grants of the Deutsche Forschungsgemeinschaft (PB) and the Ministry for Science and Culture of the Federal State of Saxony-Anhalt (AW). The experiments carried out in this study comply with the current laws of Germany and Australia. WALTER ET AL. — WEB DECORATING BEHAVIOR IN ARGIOPE KEYSERLINGI 543 LITERATURE CITED Baba, Y. & T. Miyashita. 2006. Does individual internal state affect the presence of a barrier web in Argiope bruennichi (Araneae: Araneidae)? Journal of Ethology 24:75-78. Blackledge, T.A. 1998. Stabilimentum variation and foraging success in Argiope aurantia and Argiope trifasciata (Araneae: Araneidae). Journal of Zoology 246:21-27. Blackledge, T.A. & J.W. Wenzel. 1999. Do stabilimenta in orb webs attract prey or defend spiders? Behavioral Ecology 10:372-376. Blackledge, T.A. & J.W. Wenzel. 2001. Silk mediated defense by an orb web spider against predatory mud-dauber wasps. Behaviour 138:155-171. Bonaric, J.-C. 1987. Moulting hormones. Pp. 111-118. In Ecophys- iology of Spiders. (W. Nentwig, ed.). Springer, Berlin. Bruce, M.J. 2006. Silk decorations: controversy and consensus. Journal of Zoology 269:89-97. Bruce, M.J., A.M. Heiling & M.E. Herberstein. 2005. Spider signals: are web decorations visible to birds and bees? Biology Letters 1:299-302. Bruce, M.J. & M.E. Herberstein. 2005. Web decoration polymor- phism in Argiope Audouin, 1826 (Araneidae) spiders: ontogenetic and interspecific variation. Journal of Natural History 39:3833— 3845. Bruce, M.J., M.E. Herberstein & M.A. Elgar. 2001. Signaling conflict between prey and predator attraction. Journal of Evolutionary Biology 14:786-794. Cheng, R.-C. & I-M. Tso. 2007. Signaling by decorating webs: luring prey or deterring predators. Behavioral Ecology 18:1085-1091. Craig, C.L. 2003. Spiderwebs and Silk. Tracing Evolution from Molecules to Genes to Phenotypes. Oxford University Press. New York. 230 pp. Craig, C.L. & G.D. Bernard. 1990. Insect attraction to ultraviolet- reflecting spider webs and web decorations. Ecology 71:616-623. Craig, C.L., S.G. Wolf, L.D. Davis, M.E. Hauber & J.L. Maas. 2001. Signal polymorphism in the web-decorating spider Argiope argentata is correlated with reduced survivorship and the presence of stingless bees, its primary prey. Evolution 55:986-993. Eberhard, W.G. 1990. Function and phylogeny of spider webs. Annual Review of Ecology and Systematics 21:341-372. Eberhard, W.G. 2003. Substitution of silk stabilimenta for egg sacs by Allocyclosa bifurca (Araneae: Araneidae) suggests that silk stabilimenta function as camouflage devices. Behaviour 140:847- 868. Edmunds, J. 1986. The stabilimenta of Argiope flavipalpis and Argiope trifasciata in West Africa, with a discussion of the function of stabilimenta. Pp. 61-72. In Proceedings of the Ninth Interna- tional Congress of Arachnology, Panama. (W.G. Eberhard, Y.D. Lubin & B.C. Robinson, eds.). Smithsonian Institution Press, Washington, D.C. Eisner, T. & S. Nowicki. 1983. Spider web protection through visual advertisement: role of the stabilimentum. Science 219:185-187. Elgar, M.A., J.M. Schneider & M.E. Herberstein. 2000. Female control of paternity in the sexually cannibalistic spider Argiope keyserlingi. Proceedings of the Royal Society of London B 267:2439-2443. Ewer, R. 1972. The devices in the web of the West African spider Argiope flavipalpis. Journal of Natural History 6:159-167. Foelix, R.F. 1996. Biology of Spiders. Second edition. Oxford University Press, New York. 330 pp. Hauber, M.E. 1998. Web decorations and alternative foraging tactics of the spider Argiope appensa. Ethology, Ecology & Evolution 10:47-54. Herberstein, M.E. 2000. Foraging behaviour in orb-web spiders (Araneidae): do web decorations increase prey capture success in Argiope keyserlingi Karsch, 1878? Australian Journal of Zoology 48:217-223. Herberstein, M.E., C.L. Craig, J.A. Coddington & M.A. Elgar. 2000. The functional significance of silk decorations of orb-web spiders: a critical review of the empirical evidence. Biological Reviews 75:649-669. Herberstein, M.E. & A.F. Fleisch. 2003. Effect of abiotic factors on the foraging strategy of the orb-web spider Argiope keyserlingi (Araneae: Araneidae). Animal Ecology 28:622-628. Herberstein. M.E., A.C. Gaskett, J.M. Schneider, N.G.F. Vella & M.A. Elgar. 2005. Limits to male copulation frequency: sexual cannibalism and sterility in St Andrew’s Cross spiders (Araneae, Araneidae). Ethology 111:1050-1061. Herberstein, M.E. & I-M. Tso. 2000. Evaluation of formulae to estimate the capture area and mesh height of orb webs (Araneoidea, Araneae). Journal of Arachnology 28:180-184. Higgins, L.E. 1990. Variation in foraging investment during the intermolt interval and before egg-laying in the spider Nephila clavipes (Araneae: Araneidae). Journal of Insect Behavior 3:77-783. Higgins, L.E. & M.A. Rankin. 1999. Nutritional requirements for web synthesis in the tetragnathid spider Nephila clavipes. Physio- logical Entomology 24:263-270. Higgins, L.E. & M.A. Rankin. 2001. Mortality risk of high rate of weight gain in the spider Nephila clavipes. Functional Ecology 15:24-28. Horton. C.C. 1980. A defense function for the stabilimenta of two orb weaving spiders (Araneae, Araneidae). Psyche 87:13-20. Humphreys. W.F. 1992. Stabilimenta as parasols: shade construc- tion by Neogea sp. (Araneae: Araneidae: Argiopinae) and its thermal behaviour. Bulletin of the British Arachnological Society 9:47-52. Hutchinson, J.M.C., J.M. McNamara, A.I. Houston & F. Vollrath. 1997. Dyar’s rule and the investment principle: optimal moulting strategies if feeding rate is size-dependent and growth is discontinuous. Philosophical Transactions of the Royal Society B 352:113-138. Janetos, A.C. 1982. Foraging tactics of two guilds of web-spinning spiders. Behavioral Ecology and Sociobiology 10:19-27. Kerr, A.M. 1993. Low frequency of stabilimenta in orb webs of Argiope appensa (Araneae: Araneidae) from Guam: an indirect effect of an introduced avian predator? Pacific Science 47:328-337. Li, D. 2005. Spiders that decorate their webs at higher frequency intercept more prey and grow faster. Proceedings of the Royal Society of London B 272:1753-1757. Lubin. Y.D. 1975. Stabilimenta and barrier webs in the orb webs of Argiope argentata (Araneae, Araneidae) on Daphne and Santa Cruz Islands, Galapagos. Journal of Arachnology 2:1 19-126. McCook, H.C. 1889. American Spiders and Their Spinning Work. Academy of Natural Sciences, Philadelphia. 373 pp. Nentwig, W. 1986. Stabilimente in Radnetzen von Argiope- hrien (Araneae: Araneidae): ein funktionsloses Relikt? Verhandlungen der Deutschen Zoologischen Gesellschaft 79:182. Nentwig, W. & S. Heimer. 1987. Ecological aspects of spider webs. Pp. 211-225. In Ecophysiology of Spiders. (W. Nentwig, ed.). Springer, Berlin. Olive, C.W. 1980. Foraging specializations in orb-weaving spiders. Ecology 61:1 133-1 144. Opell. B.D. 1998. Economics of spider orb-webs: the benefits of producing adhesive capture thread and of recycling silk. Func- tional Ecology 12:613-624. Peakall, D.B. 1971. Conservation of web proteins in the spider, Araneus diadematus. Journal of Experimental Zoology 176:257-264. 544 THE JOURNAL OF ARACHNOLOGY Peters, H.M. 1953. Beitriige zur vergleiclienden Ethologie und Okologie tropischer Webespinnen. Zeitschrift fur Morphologie und Okologie der Tiere 42:278-306. Peters, H.M. 1993. Uber das Problem der Stabilimente in Spinnen- netzen. Zoologische Jahrbucher, Abteilung fiir allgemeine Zoolo- gie und Physiologie der Tiere 97:245-264. Pulz, R. 1987. Thermal and water relations. Pp. 26-55. In Ecophysiology of Spiders. (W. Nentwig, ed.). Springer, Berlin. Rao, D., K. Cheng & M.E. Herberstein. 2007. A natural history of web decorations in the St Andrew’s Cross spider ( Argiope keyserlingi). Australian Journal of Zoology 55:9-14. Robinson, B.C. & M.H. Robinson. 1974. The biology of some Argiope species from New Guinea: predatory behaviour and stabilimentum construction (Araneae: Araneidae). Zoological Journal of the Linnean Society 54:145-159. Robinson, B. & M.H. Robinson. 1978. Developmental studies of Argiope argentata (Fabricius) and Argiope aemula (Walckenaer). Symposia of the Zoological Society of London 42:31-40. Robinson. M.H. & B. Robinson. 1970. The stabilimentum of the orb web spider, Argiope argentata : an improbable defence against predators. Canadian Entomologist 102:641-655. Robinson, M.H. & B. Robinson. 1973. The stabilimenta of Nephila clavipes and the origins of stabilimentum-building in araneids. Psyche 80:277-288. Scharff, N. & J.A. Coddington. 1997. A phylogenetic analysis of the orb-weaving spider family Araneidae (Arachnida, Araneae). Zoological Journal of the Linnean Society 120:355-434. Schoener, T.W. & D.A. Spiller. 1992. Stabilimenta characteristics of the spider Argiope argentata on small islands: support of the predator-defense hypothesis. Behavioral Ecology and Sociobiology 31:309-318. Seah, W.K.. & D. Li. 2002. Stabilimentum variations in Argiope versicolor (Araneae: Araneidae) from Singapore. Journal of Zoology 258:531-540. Simon, E. 1895. Histoire Naturelle des Araignees, Volume 1, fascicle 4. Libraire Encyclopedique de Roret, Paris. Pp. 761-1084. Starks. P.T. 2002. The adaptive significance of stabilimenta in orb- webs: a hierarchical approach. Annales Zoologici Fennici 39:307-315. Tanaka, K. 1984. Rate of predation by a kleptoparasitic spider, Agyrodes fissifrons, upon a large host spider, Agelena limbata. Journal of Arachnology 12:363-367. Tolbert, W.W. 1975. Predator avoidance behaviors and web defensive structures in the orb weavers Argiope aurantia and Argiope trifasciata (Araneae, Araneidae). Psyche 82:29-52. Townley, M.A. & E.K. Tillinghast. 1988. Orb web recycling in Araneus cavaticus (Araneae, Araneidae) with an emphasis on the adhesive spiral component, GABamide. Journal of Arachnology 12:303-319. Townley, M.A., E.K. Tillinghast & C.D. Neefus. 2006. Changes in composition of spider orb web sticky droplets with starvation and web removal, and synthesis of sticky droplet compounds. Journal of Experimental Biology 209:1463-1486. Tso, I-M. 1996. Stabilimentum of the garden spider Argiope trifasciata : a possible prey attractant. Animal Behaviour 52:183-191. Tso, I-M. 1999. Behavioral response of Argiope trifasciata to recent foraging gain: a manipulative study. American Midland Naturalist 141:238-246. Tso, I-M. 2004. The effect of food and silk reserve manipulation on decoration-building of Argiope aetheroides. Behaviour 141:603- 616. Vollrath, F. 1987a. Growth, foraging and reproductive success. Pp. 274-286. In Ecophysiology of Spiders. (W. Nentwig, ed.). Springer, Berlin. Vollrath, F. 1987b. Kleptobiosis in spiders. Pp. 357-370. In Ecophysiology of Spiders. (W. Nentwig, ed.). Springer, Berlin. Walter, A., P. Bliss, M.A. Elgar & R.F.A. Moritz. 2008a. Argiope bruennichi shows a drinking-like behaviour in web hub decorations (Araneae, Araneidae). Journal of Ethology. Doi 1 0. 1 007/s 1 0 1 64- 007-0077-5. Walter, A., M.A. Elgar, P. Bliss & R.F.A. Moritz. 2008b. Wrap attack activates web-decorating behavior in Argiope spiders. Behavioral Ecology, doi: 10. 1093/beheco/arn030. Watanabe, T. 2000. Web tuning of an orb-web spider, Octonoba sybotides, regulates prey-catching behaviour. Proceedings of the Royal Society of London B 267:565-569. Manuscript received 31 August 2007, revised 21 April 2008. 2008. The Journal of Arachnology 36:545-551 Ontogenetic changes in web architecture and growth rate of Tengella radiata (Araneae, Tengellidae) Gilbert Barrantes and Ruth Madrigal-Brenes1: Escuela de Biologia, Ciudad Universitaria Rodrigo Facio, Universidad de Costa Rica, San Pedro, San Jose, Costa Rica Abstract. In some spiders features of the webs of early instars may represent features of the ancestor's web. Some second instar spiderlings (first instar outside of the egg sac) of Tengella radiata (Kulczynski 1909) construct a small sheet web without any type of retreat. In subsequent instars, spiderlings construct webs that consist of a sheet with a small retreat that opens near its center. Webs gradually change as spiderlings growth and webs of 7th instar spiders are indistinguishable from those of adult females. Spiders only begin to include cribellate threads in their webs during the 7th instar. The growth of T. radiata is slow during the first three instars, but spiders’ sizes increase steadily in the subsequent stages. Legs I of adult males are longer than in females, indicating an allometric growth that occurred mainly during the last molt of males. Keywords: Spiderlings’ webs, cribellate silk, cephalothorax and leg growth Little is known about ontogenetic changes in most spiders’ webs, particularly in non-orb weavers (Eberhard 1990). The first webs constructed by newly emerged spiderlings (second instar spiderlings) in several spider families differ from the webs of adults and they tend to represent less derived stages or characters compared to the features of the webs of adult spiders (Nielsen 1931; Eberhard 1977, 1985, 1986, 1990; Robinson & Lubin 1979). Some of these differences might also be related to underdevelopment of silk glands, as is the case in Uloborus diversus Marx 1898 (Eberhard 1977). Webs constructed by young spiderlings of this species lack cribellate silk, and radial threads are more numerous than in webs of adult spiders. Another possible difference in young spider- lings’ webs may also be related with the type and size of prey that spiderlings can handle (Lubin 1986). Webs of young cribellate and non-cribellate spiderlings have been studied in Orbiculariae orb-weaving spiders (Eberhard 1977, 1985, 1986). In some of these spiders in which architecture of adult webs depart from a typical orbicular web (Eberhard 1985, 1986), the webs of newly emerged spiderlings are orbicular, indicating that their ancestors probably possessed orbicular webs. A similar pattern is showed by cribellate and non-cribellate non-orb weavers such as the psechrid Fecenia sp. (Robinson & Lubin 1979) and some theridiid species (Nielsen 1931; Szlep 1965). In this study, we focus on the cribellate spider Tengella radiata (Kulczynski 1909), a species of Tengellidae that builds funnel webs and is restricted to Costa Rica (Eberhard et al. 1993). Its web is regularly inhabited by some symbiont spiders (e.g., Philoponella sp., Mysmenopsis spp.) and plokiophilid bugs (e.g, Lipokophila spp.) (Eberhard et al. 1993). The structure and production of the cribellate threads of this spider have been described and compared with threads of other cribellate spiders (Eberhard 1988; Eberhard & Pereira 1993). The courtship and copulation have also been reported and compared with those of other spiders in related families (Barrantes 2008), and Santana et al. (1990) investigated the predation rate in the field and estimated the metabolic rate of this spider. Other than this reported work, there have been no ‘Corresponding author. E-mail: ruthymad@gmail.com published studies on the ontogenetic changes in the funnel, criballete web. or on the growth of this spider. If there are ontogenetic changes in webs, we expect there to be transitional stages between the first webs and the adult webs. We describe here the architecture of the adult web and changes in the web architecture that occur between the immature stages of Tengella radiata. We also describe the number of molts, growth, and feeding behavior of spiderlings for this spider. METHODS We observed (/? > 30) and photographed (/; = 5) webs of wild mature females of T. radiata in San Antonio, Escazu, San Jose province (= SAE; 10°56'N, 84°08'W) and used this information to describe the adult web. The egg stage period and maternal behavior were described from a female raised from eggs in captivity and maintained in a plastic box (30 x 18 X 1 1 cm) where she constructed her web. In order to rear and study numerous spiderlings we collected egg sacs from two mature female from two localities. One was from an adult female in SAE and a second from a female from La Selva Biological Station (= LSBS; 10’26'N, 83°59'W), Organization for Tropical Studies, Heredia prov- ince, Costa Rica. Each egg sac was placed on a mass of dry cotton inside a plastic container ( 13 X 13x6 cm) with a little piece of humid cotton at a corner. All containers were maintained indoors at a temperature of 1 8—20 C. From one egg sac (SAE), 16 spiderlings were separated as they emerged and each one was maintained individually in a plastic container (13 X 13x6 cm); some of these spiderlings were placed in a larger container (30 X 15 X 10 cm) when they reached their seventh instar. Webs of some of these 16 individuals were photographed and their shed cuticles measured at each stage (details described below) to describe the ontogenetic changes in the web and the growth of this spider. From a second egg sac (LSBS) we separated three groups of 10 spiderlings and each group was maintained in a container (13 X 13x6 cm). Feeding behavior was observed in these groups of spiderlings. (Spiderlings from the second egg sac were released to the wild as they molted the third time as were those remaining spiderlings from the first egg sac). We consider spiderlings that recently emerged from the egg sac as second instar individuals (Foelix 1996). 545 546 THE JOURNAL OF ARACHNOLOGY The plastic container of each of the 16 spiderlings kept individually had four small rocks, whose base and height varying from 2 to 3 cm. One rolled dry leaf, forming a cone of ca 2 cm with an opening of ca 0.5 cm in diameter, was attached to one of the rocks with masking tape; rocks were all fastened to the bottom of the container. This “rock landscape” was fashioned to offer spiderlings enough supports to construct their webs, and the rolled leaf was offered as a “natural tunnel.” Two to four webs constructed by spiderlings from 3rd to 6th instars were lightly coated with cornstarch and then photographed. The longest and widest sides of these, approximately rectangular webs were measured. Spiderlings of the photographed webs were collected and preserved in alcohol to avoid possible effects of cornstarch in the construction of subsequent webs. The web of the second instar spiderlings was sketched using photos and observations under the dissecting microscope; cornstarch adhered to threads of these webs was insufficient to allow good contrast photographs. We measured the length of tibiae and femora of legs I and IV, and the length and width of the cephalothorax in shed cuticles of the different stages of those 16 spiders maintained individually in containers, three adult males, and three adult females to estimate the spiders’ growth between stages (2nd instar to adult stage). Sample sizes were not the same for each stage because some spiderlings were collected and some shed cuticles were destroyed when we withdrew them from the web, and tibiae in all second stage cuticles collapsed and were impossible to measure. To calculate the percentage of growth between subsequent stages, we subtracted the length of a particular structure to the mean of that structure of the previous stage (Mps), then divided this difference between Mps, and multiplied this proportion by 100 (e.g., [FI2 1 — MFIi]/ MFI] * 100; FI2 1 - femur I from individual 1 of second instar, MFIr mean of femur of all individuals of instar 1). We used a digital camera (Nikon, Coolpix 4500) to photograph each femur, tibia, and cephalothorax under a dissecting micro- scope, and measured them using the software program Image Tool v. 3.0. Four 3rd instar spiderlings were observed under the dissecting microscope to check for the presence of the cribellar plate and calamistrum. During the 2nd and 3rd instars, spiderlings were offered a Drosophila fly every other day, 4th instar spiderlings were offered two Drosophila flies every other day, and spiderlings of later instars were offered one blow fly (Calliphoridae) or a moth every three days. All containers had a small piece of wet cotton for the spiderlings to drink. Feeding behavior observations were made on both solitary spiderlings and on those maintained in groups. Additional behavioral observa- tions of adults and spiderlings were obtained from adult spiders raised from eggs in captivity and complemented with the field information from SAE and LSBS. Voucher specimens of spiders from all stages were deposited in the Museo de Zoologia of the Universidad de Costa Rica, San Jose. RESULTS Adult web.— The adult web of T. radiata consists of a large, more or less triangular, horizontal sheet with a dispersed tangle above and beneath it (Fig. 1; Eberhard et al. 1993). It is usually built near a large object such as rock or a tree trunk. At the “interior,” most protected section (the apex) of the web the spider constructs a tunnel that varies between 5 and 20 cm long (it = 20). The spider rests at the mouth of the tunnel or inside it during the day. At night (// = 10) the spider usually Figure 1 . - Frontal view of an adult web of T. radiata (without cornstarch): a. Threads of the upper tangle, b. Tunnel at the “interior” section of the sheet, c. Sheet. BARRANTES & MADRIGAL-BRENES— WEB ARCHITECTURE IN TENGELLA RADI AT A 547 rests at the mouth of the tunnel, but is often observed repairing the web or producing adhesive cribellate threads that she lays on the sheet and upper tangle. Newly constructed webs frequently lack cribellate threads, but this silk accumu- lates on the sheet and upper tangle over time, until the web collapses apparently due to the weigh of debris accumulated on its sheet. Adult males (n = 3) did not build webs when placed in a large plastic container. However, they killed and fed on prey that walked nearby. Females (n — 5) placed in similar containers built a complete web. Duration of egg stage and maternal care. — One spider produced two egg sacs in captivity, the first 32 and the second 48 days after copulating. Spiderlings emerged from the first egg sac 58 days later; no spiderlings emerged from the second egg sac. The egg sacs were attached to the roof of the tunnel, as were five egg sacs observed in the field. In captivity, the spider added some small pieces of prey cuticle to the external surface of the egg sacs; egg sacs observed in the field were also covered with detritus. The female spent most of her time (except when capturing prey and feeding) hanging from the tunnel with her legs and palps surrounding the egg sac. Nearly the entire time, she contacted the egg sac with legs II and III (less frequently with legs IV), and her palps. Occasionally, she stood on the bottom of the tunnel, with one of her legs (usually one leg III) raised to contact the egg sac. When the spider produced the second egg sac, she concentrated her care on the second sac. The spider died after living for 18 months and 20 days (from emergence through adulthood). Ontogenetic changes in webs. — During the second instar, 4 of the 16 spiderlings used the tunnels (rolled leaf provided) but did not construct webs, 8 remained under a rock, and 4 constructed a web that consisted of a small, more or less rectangular, sheet (Fig. 2). At first the web consisted only of more or less horizontal threads that extended between rocks or between rocks and the container wall. The horizontal threads were about 1.5 cm above the container floor and other threads connected them to the floor and to the container wall above. These threads were part of the scaffolding that supported the Figures 2-7. — Webs of early stages of T. radiata (with cornstarch): 2. Molting web of second stage spiderlings; 3. Type I web of third stage spiderling; 4. Type II web of third stage spiderling; 5. Web of fourth stage spiderling; 6. Web of fifth stage spiderling; 7. Web of seventh stage spiderling. White arrows in Figures 3, 4, 5, and 7 show the tunnel opening. The right tip of the short white line in Figure 5 shows the spiderling inside the retreat. 548 THE JOURNAL OF ARACHNOLOGY rest of the web. During the next days the spiderlings constructed a dense sheet of very fine threads. The angle of the sheet varied from nearly horizontal to about 15° (angles were visually estimated). The complete web was constructed over the first 2 to 4 days. When the web was finished, the spiderling remained motionless, near the center, and on top of the sheet until its next molt. These webs all lacked tunnels or any other type of retreat. During this stage, which lasted 11.3 days (SD = 0.4) spiderlings did not feed; they did not react to the presence of prey on their webs. The third-stage spiderlings (// = 16) constructed two types of webs. The most common type (13 out of 16) was a small (range = 3-4 x 2-2.5 cm) dense, more or less horizontal sheet with a resting place (retreat) that was constructed nearly perpendicularly under the sheet (type 1) (Fig. 3). This retreat was a small bag ofloose silk with the exit near the center of the sheet (Fig. 3). The other web (type II) was also a flat sheet, but with the bag-like retreat constructed on the sheet forming a small roof, with a relatively dense tangle above the sheet (Fig. 4). Threads of the tangle served as support for the much finer threads of the sheet and retreat. All spiderlings in this stage captured Drosophila flies dropped on the sheet. The webs of fourth-stage spiderlings were in general larger (range = 5-6 X 3-3.5 cm) than those of the previous stage. Spiderlings expanded the sheet and retreat of the web constructed in the previous stage, and constructed a tunnel under the sheet. The closed end of the tunnel nearly reached one border of the sheet (Fig. 5), and spiderlings rested deep inside the tunnel (Fig. 5). The general shape of these webs was similar to webs of the previous instar and type I and II webs were still distinguishable. Webs of fifth and sixth stages were similar in the general features but larger compared to webs of the previous stage (Fig. 6). Seventh-stage juveniles constructed a much denser tangle above the level of the sheet. In this stage the tunnel of two webs (n = 9) was constructed on the sheet for most of its length, with its opening near the center of the sheet. The furthest end of the tunnel curved down and under the sheet, similar to a type II design ( Fig. 7). The other seven webs were similar to the adult webs, with the opening of the tunnel at one extreme of the sheet. This was the first stage in which cribellate threads were observed on the sheet and upper tangle of the web (Fig. 8), though the cribellar plate and calamistrum were already present in 3rd instar spiderlings. The web of older stages (in larger containers) was indistinguish- able from webs of adult spiders. Spiderling feeding behavior. Third and fourth instar spiderlings attacked by rushing onto the sheet to the prey and biting it. If prey was small, relative to the spiderling size, the spiderling fed on the prey without releasing it. If prey was large, it was released by the spiderling as soon as the prey’s struggling stopped, returning to feed on it a few seconds later. One fourth instar spiderling (in a group of 10) wrapped the prey after biting it. The spiderling released the prey as it stopped struggling and nearly immediately began to wrap it. During wrapping, the spiderling turned in place while attaching wrapping lines to the substrate (the sheet). The wrapping movements were similar to those described for adult spiders (Barrantes & Eberhard 2007). Another spiderling approached a large blow fly (larger than a house fly) slowly. The spiderling backed away as the fly struggled and then cut Figure 8. — Cribellate threads (white arrow) from webs of seventh stage spiderlings. some sheet threads and walked, hanging under the sheet to the prey and bit one of the fly’s legs through the sheet. Neither wrapping nor attacks from under the sheet were observed in any fifth instar individuals. Older juvenile spiders (sixth or older instars) always moved on the upper surface of the sheet, and often wrapped large prey (e.g., moths and large flies), with movements similar to those of adult spiders (Barrantes & Eberhard 2007). Large prey were carried inside the tunnel by 7th stage, subadult, and adult spiders, and their carcasses were left inside the tunnel. Additional behavioral observations. — Early instar spiders did not molt at any consistent site in the web (n > 50). Flowever, older stages (7th to pre-adult) molted inside the tunnel but they carried the shed skin to the farthest extreme of the web (n = 9). Number of molts and growth. — The males had eight molts (» = 4) and the females nine (« = 4) to attain their adult stages. The mean time from emergence from the egg to the eighth molt was 186.7 days (± 10.5), when males molted to the adult stage. Females lasted 42 days (± 7) more to their next and last molt (Fig. 9). The general pattern of growth of the cephalothorax (width and length) and legs I and IV (tibia and femur) were similar. In the three first stages these structures grew very little, but during the following stages the size of the cephalothorax and legs increased steadily (Figs. 10-12). In fact, the growth of all morphological features was proportionally much larger between the third and the fourth stage, than between other subsequent stages (Table 1 ). It is also evident that the length of the cephalothorax increased faster than its width ( Fig. 1 0), indicating an allometric growth of the length relative to the width of the cephalothorax. In addition, legs of adult males were also notably longer than those of adult females, despite the additional molt of females. DISCUSSION The architecture of the webs of T. radiata changes as spiders mature. The first web constructed by 2nd instar spiderlings, which have recently emerged from the egg sac, apparently serves as a molting place, since spiderlings in this stage did not capture prey. In nature and in captivity the second instar spiderlings construct a communal molting web inside the BARRANTES & MADRIGAL-BRENES— WEB ARCHITECTURE IN TENGELLA RADIATA 549 9 55 50 45 40 35 m m 30 Q 25 20 15 10 5 123456789 Stages 11 10 Molt Figures 9-12. — Inter-molt time and growth of T. radiata : 9. Intermolt time (mean, standard error and standard deviation) from the first to the ninth stage out of the egg sac; sample size above each stage mean; 10. Mean and standard deviation of the length (solid line) and width (dotted line) of cephalothorax (cph); sample size above each mean, F and M indicate the values for adult females and adult males respectively; 1 I . Mean and standard deviation of the length of femur (solid line) and tibia (dotted line) of leg I; 12. Mean and standard deviation of the length of femur (solid line) and tibia (dotted line) of leg IV. tunnel of their mother’s web, and only begin to abandon the tunnel and the web after they molt to the third instar (GB unpublished data). Both of the two types of webs constructed by third through sixth instar spiderlings had retreats that opened near the center of the sheet (Figs. 3-7). If, as in other groups, the webs of early instars represent less derived characters than those of the adult web (Eberhard 1986, 1990), the ancestral web of Tengella might have been a sheet with a tunnel retreat extending below its center. However, comparative data of closely related species (Codding- ton 2005) are required to test this hypothesis. The lack of cribellate threads on the sheet of juvenile spiders (2nd to 6th instar) may either represent an ancestral condition, an undeveloped condition of the cribellum Table 1. — Percentage of growth of the width and length of the cephalothorax (cph) and the femur (F) and tibia (T) of legs 1 and IV between successive stages (first row). Sample size in parentheses beside or under the stage codes. VII to VIII VIII to IX I to II (5) II to III (9) III to IV (6) IV to V (9) V to VI (9) VI to VII (8) (7) (6) VIII to 3 (3) IX to 9 (3) Cph W 0.6 ± 9.9 19.9 ± 1.9 104.4 ± 11.7 30.7 ± 2.9 19.8 ± 4.3 35.3 ± 8.7 23.9 ± 5.8 18.6 ± 3.5 32.9 ± 4.7 30.5 ± 3.3 Cph L 0.1 ± 0.2 21.6 ± 3.3 106.4 ± 9.2 35.6 ± 4.4 20.1 ± 3.7 36.6 ± 9.7 26.5 ± 6.4 19.3 ± 3.4 30.5 ± 3.2 25.3 ± 2.6 FI 21.4 ± 5.2 41.0 ± 3.9 104.4 ± 14.6 46.5 ± 5.5 22.4 ± 9.8 42.0 ± 8.4 23.8 ± 8.7 23.7 ± 3.7 82.4 ± 5.9 15.8 ± 6.5 TbI 56.6 ± 5,6 114.5 ± 8.0 41.8 ± 7.3 24.4 ± 8.1 38.9 ± 5.9 30.6 ± 10.8 18.5 ± 4.5 84.8 ± 7.7 18.3 ± 7.1 FIV 5.9 ± 3.3 32.3 ± 4.0 110.9 ± 17.2 42.0 ± 5.0 26.2 ± 5.5 38.7 ± 6.0 26.3 ± 13,3 22.8 ± 5.3 67.8 ± 2.6 14.2 ± 1.4 TbIV 29.6 ± 5.5 109.5 ± 13.3 44.0 ± 4.2 25.1 ± 5.5 37.5 ± 8.3 26.3 ±11.1 23.1 ± 4.7 86.3 ± 4.3 20.2 ± 2.0 550 THE JOURNAL OF ARACHNOLOGY apparatus, or both. The presence of the cribellum and calamistrum in third instar spiderlings but the lack of criballate silk on their webs, indicate that these structures may not be functional in the early stages of T. radiata, as occur in some uloborids (Eberhard 1977; Opell 1982), or that the high demand of energy involved in drawing the cribellar fibrils (Eberhard 1988) prevent young spiderlings from using this type of silk. The lack of cribellate silk in early stages and newly constructed webs of adult T. radiata did not prevent spiders from using their webs to capture prey, though cribellate silk possibly restrain and reduce movements of prey on the web. Spiderlings at early stages showed attack and wrapping behaviors similar to those of adult spiders (Barrantes & Eberhard 2007). The wrapping behavior of spiderlings only differed from that of adults in that spiderlings did not hold the prey while wrapping it as adults do (Barrantes & Eberhard 2007). In addition, one spiderling attacked a large prey from under the sheet. This attack is apparently restricted to early spiderling stages, as it has not been observed in either large juveniles or adult spiders. The low frequency of occurrence of this behavior is unclear since attacking a large prey from under the sheet provides some protection to the spiderling as the sheet restrains the struggling prey and reduces the force of its movements (Robinson 1975; Lubin 1980). In general the growth of T. radiata is slow, as expected from its apparently extremely low metabolic rate (Santana et al. 1990). Growth was very slow during the earliest stages (Figs. 10-12), but the relative increment in cephalothorax and leg size was more than 100% between the third and fourth stage. After the fourth stage, the cephalothorax and legs increased steadily with each subsequent molt, though the intermolt period tended to increase with each stage. This possibly indicates the need for more energy and time for growth and development of internal organs as spiders mature. Length of legs 1 and IV is nearly the same until the fourth instar. However, in the following stages, increments in the length of leg I are larger than in leg IV, possibly reflecting the different functions of these legs in young and adult spiders. For example, the tactile function of legs I likely favor their longer length (Foelix 1996). The comparatively longer legs of adult males result from the allometric growth that occurred mainly during the last molt of males, as happens in wolf spiders (Framenau 2005). Though experimental evidence is lacking, field observations (e.g., males observed near or on females’ webs) indicate that adult males abandon their webs to find receptive females, so that longer legs may result in greater step size to bridge gaps. If a more efficient search for females lead to a higher reproductive success in males, it is likely that longer legs in males evolved, at least partially, through indirect male-male sexual competition (Anderson 1994; Framenau 2005). Additionally, natural selection might have also favored longer legs in males, if such a trait allows them to run faster to escape from predators and from females during courtship, and provides them a greater sensory range (Gertsch 1949; Framenau 2005). Accordingly, the larger size of adult females’ cephalothorax correlates with their larger body which is related to their capacity to produce large numbers of eggs (Gertsch 1949). ACKNOWLEDGMENTS We thank W.G. Eberhard, G. Stratton, and two anonymous reviewers for providing valuable comments that greatly improved this manuscript. We also thank J-L Weng for the picture of cribellate threads. Research was supported by Vicerrectoria de Investigation, Universidad de Costa Rica. LITERATURE CITED Anderson, M. 1994. Sexual Selection. Princeton University Press, Princeton, New Jersey. 599 pp. Barrantes, G. 2008. Courtship behavior and copulation in Tengella radiata (Araneae, Tengellidae). Journal of Arachnology 36:606-608. Barrantes, G. & W.G. Eberhard. 2007. The evolution of prey wrapping behaviour in spiders. Journal of Natural History 41:1631-1658. Coddington, J.A. 2005. Phylogeny and classification of spiders. Pp. 18-24. In Spiders of North America: an identification manual. (D. Ubick, P. Paquin, P.E. Cushing & V. Roth, eds.). American Arachnological Society. Eberhard, W.G. 1977. The webs of newly emerged Uloborus diversus and of male Uloborus sp. (Araneae, Uloboridae). Journal of Arachnology 4:201-206. Eberhard, W.G. 1985. The “sawtoothed” orb of Eustala sp., with a discussion of the ontogenetic patterns of change in web design in spiders. Pyche 92:105-118. Eberhard, W.G. 1986. Ontogenetic changes in the web of Epeirotypus sp. (Araneae, Theridiosomatidae). Journal of Arachnology 14:125-128. Eberhard, W.G. 1988. Combing and sticky attachment behaviour by cribellate spiders and its taxonomic implications. Bulletin of the British Arachnological Society 7:247-251. Eberhard, W.G. 1990. Function and phylogeny of spider webs. Annual Review of Ecology and Systematics 21:341-372. Eberhard, W.G. & F. Pereira. 1993. Ultrastructure of cribellate silk of nine species in eight families and possible taxonomic implications (Araneae: Amaurobiidae, Deinopidae, Desidae, Dictynidae. Filis- tatidae, Hypochilidae, Stiphidiidae, Tengellidae). Bulletin of the British Arachnological Society 21:161-174. Eberhard, W. G, N. Platnick & R.T. Schuh. 1993. Natural history and systematics of arthropod symbionts (Araneae; Hemiptera; Diptera) inhabiting webs of the spider Tengella radiata (Araneae, Tengellidae). American Museum Novitates No. 3065. 17 pp. Foelix, R.F. 1996. Biology of Spiders. Second edition. Oxford University Press, New York. 330 pp. Framenau, V.W. 2005. Gender specific differences in activity and home range reflect morphological dimorphism in wolf spiders (Aranea, Lycosidae). Journal of Arachnology 33:334-346. Gertsch, W.J. 1949. American Spiders. D. Van Nostrand Company, Inc., Princeton, New Jersey. 285 pp. Lubin, Y.D. 1980. The predatory behavior of Cyrtophora (Araneae, Araneidae). Journal of Arachnology 8:159-183. Lubin, Y.D. 1986. Web building and prey capture in the Uloboridae. Pp. 132-171. In Spiders: Webs, Behavior, and Evolution. (W.A. Shear, ed.). Stanford University Press, Stanford, California. Nielsen, E. 1931. The Biology of Spiders with Especial Reference to the Danish Fauna. Levin & Munksgaard, Copenhagen, Denmark. 248 pp. Opell, B.D. 1982. Post-hatching development and web production of Hyptiotes cavatus (Hentz) (Araneae, Uloboridae). Journal of Arachnology 10:185-191. Robinson, M.H. 1975. The evolution of predatory behaviour in araneid spiders. Pp. 292-312. In Function and Evolution in Behaviour. (G. Baerends, C. Beer & A. Manning, eds.). Clarendon Press, Oxford, UK. Robinson, M.H. & Y.D. Lubin. 1979. Specialists and generalists: the ecology and behavior of some web-building spiders from Papua BARRANTES & MADRIGAL-BRENES— WEB ARCHITECTURE IN TENGELLA RADI AT A 551 New Guinea II. Psechrus argentatus and Fecenia sp. (Araneae: Psechridae). Pacific Insects 21:133-164. Santana, M., W.G. Eberhard, G. Bassey, K.N. Prestwich & R.D. Briceno. 1990. Low predation rates in the field by the tropical spider Tengella radiata (Araneae: Tengeilidae). Biotropica 22:305-309. Szlep, R. 1965. The web-spinning process and web-structure of Latrodectus tredecimguttatus, L. pallidus and L. revivensis. Proceedings of the Zoological Society of London 145:75— 89. Manuscript received 24 September 2007, revised 26 May 2008. 2008. The Journal of Arachnology 36:552-556 Does the microarchitecture of Mexican dry forest foliage influence spider distribution? Pablo Corcuera1, Maria Luisa Jimenez2 and Pedro Luis Vaiverde1: 'Universidad Autonoma Metropolitana-Iztapalapa, Departamento de Biologia, Av. San Rafael Atlixco 186, Col. Vicentina, Iztapalapa, C.P. 09340, Mexico D.F.; 2Centro de Investigaciones Biologicas del Noroeste, Apdo. Postal 128, La Paz, B.C.S. 23090, Mexico. E-mail: pcmr@xan um . uam . mx Abstract. Spider species diversity has been associated with vegetation structure and stratification but there are few studies comparing the spider distribution in different shrubs and trees. In this study we analyzed the species distribution of the spider community of 1 1 shrub and tree species in two different study sites in a Mexican tropical dry forest. We present results from multivariate analyses that explain their distribution. A classification analysis based on spider abundances separated one shrub, Croton ciliatoglanduliferus , from the rest of the plant species. This was explained by the presence of large numbers of the oxyopid Peucetia viridans (Hentz 1832) on this plant. A second cluster segregated broad-leaved from small-leaved, bipinnate species. This was mainly due to higher spider abundances in the latter type of plants. Four vegetation variables were estimated and their influence on the species distribution was assessed by means of a principal components and regression analysis. With the exception of P. viridans, all spiders were positively associated with number of leaves and number of branchlets per 50 cm branch and negatively with foliage area. Keywords; Community ecology, species abundances, plant structure Habitat structure is an important factor that influences diversity, abundance, and distribution of spider species (Lubin 1978; Hatley & MacMahon 1980; Evans 1997; Whitmore et al. 2002). The available evidence has been gathered from both natural communities (e.g., Lubin 1978; Robinson 1981; Raizer & Amaral 2001) and agricultural systems (Rypstra et al. 1999; Samu et al. 1999). Habitat structure and complexity are related to factors such as prey abundance, shelter against enemies and suitable microclimatic conditions (Riechert & Tracy 1975; Gunnarsson 1996; Halaj et al. 1998; Raizer & Amaral 2001). Habitat preferences, however, can be highly specific and species belonging to different guilds have particular requirements according to their morphological, physiological, and behavioral features (Turnbull 1973; Wise 1993). Variation in plant height, foliage density, leaf surface area, number of leaves and branchlets, and number and type of inflorescences, can affect the abundance and distribution of foliage-dwelling spiders (Hatley & MacMahon 1980; Evans 1997; Halaj et al. 1998; Uetz et al. 1999; Raizer & Amaral 2001; Corcuera et al. 2004; Heikkinen & MacMahon 2004; Souza & Martins 2004, 2005). In this study, we evaluated the influence of plant architecture on the spider community by means of multivariate and regression analyses. We analyzed the abundance of foliage spiders and four plant attributes of 1 1 of the most abundant trees and shrubs found in a tropical dry forest in western Mexico. Information on Mexican spiders is widely dispersed. After extensive bibliographical research, Jimenez (1996) found 7,916 species. It is not known how many specimens were collected from foliage since most studies were concerned with taxonomy (Jimenez 1996). There have been a few reports on foliage spiders on cacao and coffee plantations (Ibarra Nunez et al. 1995, 1997; Moreno-Molina et al. 2001; Pinkus-Rendon et al. 2006), but these studies concentrate on species richness and diversity. Besides a diversity analysis (Corcuera et al. 2004), to our knowledge nothing has been written on the distribution of foliage spider communities in dry forests. METHODS Study sites. — Tropical dry forests cover 42% of the tropical and subtropical land area on the planet (Murphy & Lugo 1986). The dominant plant species are strongly drought deciduous (Mooney et ai. 1989). In Mexico, they are the prevailing vegetation type along the west coast and cover ca. 12.4% of the country’s area (Arizmendi et al. 1990). Mexican tropical dry forests are found between 0 and 1990 m elevation (Rzedowski 1978). Mean annual temperature ranges from 20 to 29°C, and mean annual precipitation from 300 to 1800 mm (Rzedowski 1978). Dry forests are strongly seasonal, with a long dry season and intense rainy season (Rzedowski 1978; Murphy & Lugo 1986). The study sites are located in the Municipality of Villa Corona in the state of Jalisco (20°20'N, 103°35'W). Altitude above sea level is 1 640 m. Mean annual temperature was 20.3°C and mean annual precipitation from the last 15 years was 826 mm. Most of the rain falls between mid-June to mid- September and there are between 6 and 8 dry months each year. Plant variables. — Eleven trees and shrubs were sampled in two sites (El Caracol and Charco Verde) to test the effect of plant architecture on the distribution of foliage spiders: Bursera schlechtendalii, B. bipinnata, Guazuma ulmifolia , Heliocarpus appendiculatus , Ipomoea woleottiana, Prosopis juliflora. Mimosa galeotti, Lysiloma acapulcense , Croton ciliatoglanduliferus , Acacia cymbispina, and Byrsonima sp. (Table 1). These plant species are typical of Mexican dry and thorn forests and were the most common shrubs and trees in the study sites (Table 2). Details about plant cover estimation are given elsewhere (Corcuera & Butterfield 1999; Corcuera & Zavala-Hurtado 2006). Foliage area, number of leaves and number of branchlets (i.e., small branches) were determined for each plant species. The sample unit was a 50 cm terminal branch from a limb rising horizontally from the center of the plant (McCaffrey et al. 1984). Foliage area was measured by drawing the contour 552 CORCUERA ET AL.— PLANT ARCHITECTURE AND SPIDER DISTRIBUTION 553 Table 1. — Mean (± SD, n = 10 per species) of plant height, foliage area, number of leaves and branchlets on a 50 cm terminal branch in = 3 per species) for 1 1 species in two dry forest sites, El Caracol and Charco Verde, in the Municipality of Villa Corona, Jalisco, Mexico. * = small- leaved species. Plant species Code Plant height (m) Foliage area (cm2) Number of leaves Number of branchlets Bursera schlechtendalii Busc 3.4 (0.32) 850 (386.1) 70.0 (28.28) 22.5 (6.4) Bursera bipinnata * Bubi 3.4 (0.71) 541 (140.3) 2473.2 (2692.16) 22.0 (9.9) Croton ciliatoglanduliferus Crci 1.1 (0.32) 1026 (631.8) 28.0 (16.17) 9.0 (4.2) Guazuma ulmifolia Guul 4.8 (0.98) 3349 (989.7) 45.8 (19.30) 21.5 (0.7) Acacia cymbispina * Accy 3.3 (0.71) 120 (27.6) 10670.4 (2313.43) 10.5 (4.9) Prosopis juliflora * Prju 3.9 (1.25) 242 (105.1) 6696.0 (2136.01) 11.0 (0.7) Byrsonima sp. Bysp 2.7 (0.70) 1712 (1065.1) 64.5 (21.71) 15.0 (4.2) Ipomoea wolcottiana Ipwo 5.0 (0.72) 2556 (960.2) 12.2 (7.79) 11.0 (2.1) Heliocarpus appendiculatus Heap 5.3 (0.67) 2050 (975.0) 15.0 (7.53) 8.0 (1.4) Lysiloma acapulcense * Lysp 4.4 (0.47) 340 (105.9) 30240.0 (11671.35) 11.5 (3.5) Mimosa galeotti * Miga 2.7 (0.48) 320 (209.6) 16301.0 (5250.44) 11.5 (7.8) of all leaves present on the branch on millimetric paper with 1 mm divisions. The procedure was repeated on three branches for each species and the mean area (cm2) per branch was calculated. Mean number of leaves, or leaflets for bipinnate species, and branchlets per branch was obtained from the three samples of each species. Plant height was recorded in a sample of 10 individuals for each species. All plant variables were averaged from measures from both sites. Spiders. — Spiders were collected by branch beating (South- wood 1978) in June, July, September, October, and November 1999, and January and April 2000. For each plant species, a terminal branch was chosen and beaten 10 times with a cane (trial samplings showed that more strokes did not dislodge more specimens) (Southwood 1978; McCaffrey et al. 1984). This procedure was repeated on 10 individuals of each plant species in each of the seven visits to each site. The specimens were collected in 60 cm diameter muslin covered trays. Two persons collected the spiders from the canvas using tweezers and manual aspirators. McCaffrey et al. (1984) found that this technique efficiently sampled the arachnofauna of foliage dwelling spiders. The number of individuals for each spider species (11 plants X 7 dates) was added in order to execute the analyses, and the results were log transformed to obtain a normal distribution. The specimens were preserved in 70% alcohol and identified later at the Centro de Investigaciones Biologicas (CIBNOR) in La Paz, Baja California. Voucher specimens have been deposited in the collection at the Laboratorio de Ecologia Animal, UAM-Iztapalapa, Mexico City. Multivariate analyses, — We analyzed the spider community similarities with a classification using an unweighted pair group average method with percent similarity. A Principal Components Analysis (PCA) was used to analyze the distribution of spider species in relation to the plant species. Regressions were used to assess the relationship between the main PCA axes and the plant variables. The classification and ordination analyses were carried out using the statistical software MVSP 3.2 (Multivariate Statistical Package; Kovach 1999). RESULTS Plant variables. — Foliage area was greater for broad-leaved trees. G. ulmifolia had the largest area (3349 cm-), followed by I. wolcottiana (2556 cm2), and H. appendiculatus (2050 cm2). Croton ciliatoglanduliferus and Byrsonima sp. had intermediate foliage areas (1026 cm2 and 1712 cm2, respectively); both are broad-leaved shrubs. All small-leaved species had much lower foliage areas (Table 1). Lysiloma acapulcense and M. galleotti, small-leaved species, had the highest mean number of leaves per terminal branch Table 2. — Plant cover percentage and spider abundance and richness of 1 1 trees and shrub species in two dry forest sites, El Caracol (C) and Charco Verde (V), in the Municipality of Villa Corona, Jalisco, Mexico. * = small-leaved species. Species Plant cover (%) Spider abundance Spider richness C V C V C V Bursera schlechtendalii 0.6 2.7 49 31 9 9 Bursera bipinnata * 2.9 3.3 68 84 11 11 Croton ciliatoglanduliferus 9.9 0.8 87 83 11 6 Guazuma ulmifolia 2.3 5.4 28 29 7 1 1 Acacia cymbispina * 31.1 23.2 128 108 12 9 Prosopis juliflora * 4.8 0.5 70 119 10 11 Byrsonima sp. 4.8 1.6 43 27 12 7 Ipomoea wolcottiana 16.0 6.0 35 35 8 8 Heliocarpus appendiculatus 1.9 7.9 32 45 10 6 Lysiloma acapulcense * 2.9 18.8 53 65 10 11 Mimosa galeotti * 11.6 4.9 74 47 1 1 10 Other plant species 11.1 24.8 — — — — Total 100 100 667 673 — — 554 THE JOURNAL OF ARACHNOLOGY r 4 — r~ 36 — r~ 84 CrciV CrciC MigaV *' GuulV BuscV ByspV HeapC IpwoC IpwoV ByspC GuulC BubiV * PrjuV ® AccyV* AccyC* HeapV PrjuC * LyspV « MigaC «■ BubiC * LyspC * BuscC . 20 52 68 100 Percent Similarity Figure 1. — Classification of 1 1 plant species in two sites according to spider species abundance. Plant codes are the same as in Table 1. The additional last capital letter represents the sample site (C = Caracol, V = Charco Verde). * = small-leaved species. (30,240 and 16,301, respectively), while the broad leaved trees, H. appendiculatus and I. wolcottiana, had the lowest number of leaves (15 and 12.2, respectively) (Table 1). Mean number of branchlets per terminal branch was higher for B. schlechten- dalii (22.5), Bursera bipinnata (22), and G. ulmifolia (21.5), while C. ciliatoglanduliferus (9) and H. appendiculatus (8) had the lowest values (Table 1). Mean plant height among species varied from 1.1m (C. ciliatoglanduliferus ) to 5.3 m (H. appendiculatus) (Table 1). The dominant species in both sites was A. cymbispina, a shrub that grows in areas that have been altered by cattle and goat grazing. In both sites, P. juliflora was the second most abundant species. C. ciliatoglanduliferus is an invasive shrub particularly common in one site (El Caracol). In this site M. galeottii was also dominant, while L. acapulcense was common in Charco Verde (Table 2). Spider abundance and composition. — A total of 1340 adult spiders belonging to 2 1 species were caught in the two sampled sites (667 in El Caracol, and 673 in Charco Verde) (Table 2). Species composition was similar in both sites. Isaloides cf yollotl (Jimenez, 1992), Hamataliwa puta (O. Pickard-Cam- bridge 1894) and Peucetia viridans (Hentz 1832) represented 73% of the total numbers caught in El Caracol and 69% in Charco Verde. Four species were represented by only one individual: Micrathena gracilis (Walckenaer 1805) and Mallos sp. in El Caracol, and Euryopis calif arnica Banks 1904 and Ocrepeira sp. in Charco Verde. The other species found were Neoscona oaxacensis (Keyserling 1864), Euriophora edax (Blackwall 1863), Wamba crispulum (Simon 1895), Theridion sp., Mimetus puritans Chamberlin 1923, Tmarus ehecaltocatl Jimenez 1992, Misumenoides sp., Modysticus cf. floridana (Banks 1895), Apollophanes punctipes (O.P. -Cambridge 1891), Philodromus albicans O. Pickard-Cambridge 1897, Oxyopes bifidus F.O. Pickard-Cambridge 1902, Phidippus sp., Para- marpissa pi" a tic a. (Peckham & Pekham 1888) and Metaphi- dippus cf. apicalis F.O. Pickard-Cambridge 1901. In both study sites, the dominant families were hunters, in particular Oxyopidae with 49% and 43% (El Caracol and Charco Verde, respectively), followed by Thomisidae (36% and 38%). The family Salticidae was represented by 10% and 1 1% of the total catch. In spite of spiders being sampled from the foliage, web weavers were only represented by 1.5% of the total catch in El Caracol, and 5.1% in Charco Verde. Spider species distribution, — A classification of the plants based on the spider abundances resulted in three main clusters at the 50% similarity level (Fig. 1). The first cluster separated C. ciliatoglanduliferus (Crci) from all other plant species. The second cluster included all the broad-leaved species with the exception of M. galeotti (MigaV) from Charco Verde, while the third cluster included all the small-leaved plants and two broad-leaved trees, H. appendiculatus (HeapV) from Charco Verde and Bursera schlechtendatii (BuscC) from El Caracol. The first division was explained by the presence of P. viridans, one of the most abundant species, which was found almost exclusively on C. ciliatoglanduliferus . Most spiders had higher abundances in small-leaved plants (Table 2), which explains the separation between the second and third clusters (Fig. 1). The first two PCA axes based on spider abundances accounted for 88% of the variance (58% and 30%, respective- ly). Since some spider species had less than 5 individuals, only 15 out of 21 species were included in the analysis. Nine of these species were common to both sites. The first axis of the ordination (eigenvalue = 0.58) was negatively correlated with plant height (r = —0.87, P < 0.001). The ordination along this axis was determined by the large numbers of P. viridans on C. ciliatoglanduliferus (both had the highest scores on the positive side (Fig. 2). The second axis (eigenvalue = 0.55) was negatively correlated with foliage area (r = —0.91, P < 0.001) and positively with number of leaves (r = 0.86, P < 0.001) and branchlets (r = 0.61, P < 0.05). The ordination pattern along this axis clearly segregated all spider species and small-leaved bipinnate plants from broad-leaved plant species (Fig. 2). Spider species, as well as all small-leaved bipinnate species had positive scores. These plants had a high numbers of leaves and branchlets, and low foliage area (Table 1, CORCUERA ET AL.- -PLANT ARCHITECTURE AND SPIDER DISTRIBUTION 555 0.6 lyoV Axis I HpuV PhspV Lysp Accy PpiC HpuC lylpuV Prju lyoC FhspC Thi? MspC PptfMiga ThspV feuC jrehV Mspv' Bubi -0.4 Bysp Heap Busc Guui ipwo -0.2 Axis 2 PviC PviV 0.8 Crci Figure 2. — Principal Components Analysis of the spider species present in 1 1 dry forest plant species. Spiders are: Pvi = Peucetia viridans, Iyo = Isaloides cf yollotl, Hpu = Hamataliwa puta , Mpu = Mimetus puritans , Phsp = Phiddipus sp., Ppi = Paramarpissa piratica , The = Truants ehecatlocatl, Thsp = Theridon sp.. Misp = Misumenoides sp. In the codes for spider species names, the additional last capital letter represents the sample site (C = Caracol, V = Charco Verde). Codes for plant names are in bold and are the same as in Table 1. Fig. 2). Conversely, the plant species on the negative side included those with high foliage area but low number of leaves and branchlets (Table 1, Fig. 2). The ordination shows the relationship of each spider species with the plants. For instance, I. cf. yollotl , a common spider species was particularly abundant on A. cymbispina and that is why these species appear together in Fig. 2. DISCUSSION In spite of intensive sampling (a total of 1540 branches during a seven month period), only 21 species were found among 1340 individuals collected in the two sampled sites (677 in El Caracol, and 673 in Charco Verde). Rarefaction analyses (P. Corcuera, unpublished results) showed that only two or three additional foliage species are likely to be found in the study area. Small-leaved plants appear to be suitable sites for foliage spiders. Evans (1997) found that social crab spiders preferred Eucalyptus species with smaller leaves. Perhaps more important- ly, and regardless of plant taxon, density of leaves per branch (e.g., Gunnarsson 1990; Souza & Martins 2005) as well as structural complexity are better predictors of spider diversity. Branching or twig density, as well as leaf density have been found to be strongly related with number of spiders, diversity, and abundance of various functional groups (Hatley & MacMahon 1980; Halaj et al. 1998; Corcuera et al. 2004). These variables also explained the spider distribution in this study. A classification of the plant species (Fig. 1) separated most small-leaved bipinnate trees and shrubs from most broad- leaved species. The second axis of an ordination also segregated the plants and gave high positive scores to all spiders and small-leaved plants and negative to all broad- leaved (Fig. 2). This axis was positively correlated with number of leaves and branchlets and negatively with foliage area. The first axis was correlated with plant height and was explained by high numbers of Peucetia viridans , a very common spider in the study sites and the only one that was associated with the small shrub Croton ciliatoglanduliferus. Causal explanations for habitat preferences of foliage spiders have not been explored in depth but some hypothesis have been suggested. For example, Peucetia species are known to favor plants with glandular trichomes, presumably because arthropods are trapped by these hairs and represent available prey for the spider (Vasconcelos-Neto et al. 2006). Halaj et al. (1998) suggested that plants with higher cover are easier to locate and might provide more resources. This might explain higher diversities of most spiders on the most common trees and shrubs. Total number of individuals was significantly correlated with plant cover in the two sites (r — 0.83, P < 0.005 for El Caracol and r = 0.67, P < 0.05 for Charco Verde). This may explain why A. cymbispina , which had the highest plant species cover in both sites, supported high densities of most spiders (Table 2). In the same way, M. galeotti , a small-leaved tree from Chaco Verde, was included in the broad-leaved cluster in the classification (Fig. 1, Table 2). This species had low densities of spiders probably because of its low cover in this site. However, P. juliflora and B. bipinnata , with high richness and species abundances, had very small cover in one or both sites (Table 2). Plants with higher cover might be easier to locate, but they do not necessarily provide more resources. Other factors (i.e., branch and leaf density) appear to be more important to understand the distribution of foliage spiders. Some plant attributes might provide suitable microclimatic conditions. Riechert & Tracy (1975) suggested that certain plants might be favored because of their ability to modify the thermal environment. In hot climates with long drought periods, preserving body temperature would be a most important factor. Small-leaved plants, especially C. cymbispina and P. juliflora could provide a cooler environment because they either do not shed their leaves (as does P. juliflora) or remain green during the early draught, which is when 556 THE JOURNAL OF ARACHNOLOGY spiderlings start to disperse. This species also starts producing leaves early, before the rains, when broad-leaved trees and shrubs are still deciduous. Once settled on these plants, there would be no reason to move to shrubs or trees where conditions would be less favorable. Besides resource availability and favorable physical condi- tions, accessibility of refuges against predators plays an important role in determining spider distribution. Gunnarsson (1996) suggested that high leaf densities could provide shelter from bird predation. This would not seem the case in our study sites, since bird attacks tended to be more frequent in small-leaved trees and shrubs (Corcuera 2001), where spiders are more abundant. Few studies have compared differences in the abundance of spiders on foliage of different shrub and tree species (e.g., Halaj et al. 1998; Raizer & Amaral 2001; Souza & Martins 2004). Although some spider species were found in small numbers (< 5 individuals), and it is not possible to reach any conclusions about their distribution, our results reveal that foliage spider species were positively influenced by small-leaved trees and shrubs with a high number of leaves and branches, and negatively by broad-leaved plants with a high foliage area among 1 1 plant species of the Mexican tropical dry forest. LITERATURE CITED Arizmendi, M.C., H. Berlanga, L.M. Marquez-Valdelamar, L. Navarijo & J.E. Ornelas. 1990. Avifauna de la Region de Chamela, Jalisco. Cuaderno No. 4, Instituto de Biologia. UNAM, Mexico. 62 pp. Corcuera, P. 2001. Plant use and the abundance of four bird guilds in a Mexican dry forest-oak woodland gradient in two contrasting seasons. Huitzil 2:2-14 (http://www.huitzil.net/ci-02pdf.pdf). Corcuera, P. & J.E.L. Butterfield. 1999. Bird communities of dry forest and oak woodland of western Mexico. Ibis 141:240-255. Corcuera, P., M.L. Jimenez & G. Lopez. 2004. Comparacion de la diversidad de aranas asociadas al follaje en una selva baja caducifolia de Jalisco. Contactos Octubre-Diciembre: 17-26. Corcuera, P. & J.A. Zavala-Hurtado. 2006. The influence of vegetation on bird distribution in dry forests and oak woodlands of western Mexico. Revista de Biologia Tropical 54:657-672. Evans, T.A. 1997. Distribution of social crab spiders in eucalypt forests. Australian Journal of Ecology 22:107-1 1 1. Gunnarsson, B. 1990. Vegetation structure and the abundant and size distribution of spruce-living spiders. Journal of Animal Ecology 59:743-752. Gunnarsson, B. 1996. Bird predation and vegetation structure affecting spruce-living arthropods in a temperate forest. Journal of Animal Ecology 65:389-397. Halaj, J.. D.W. Ross & A.R. Moldenke. 1998. Habitat structure and prey availability as predictors of the abundance and community organization of spiders in western Oregon forest canopies. Journal of Arachnology 26:203-220. Hatley, C.L. & J.A. MacMahon. 1980. Spider community organiza- tion: seasonal variation and the role of vegetation architecture. Environmental Entomology 9:632-639. Heikkinen, M.W. & J.A. MacMahon. 2004. Assemblages of spiders on models of semi-arid shrubs. Journal of Arachnology 32:313-323. Ibarra Nunez, G., A. Garcia & M. Moreno. 1995. La comunidad de artropodos de dos cafetales con diferentes practicas agricolas (organico y convencional): el caso de las Aranas. Resumenes XXX Congreso Nacional de la Sociedad Mexicana de Entomologia, Texcoco, Edomex, Mexico. Pp. 12-13. Ibarra Nunez, G., A. Garcia & M. Moreno. 1997. Diversidad de aranas tejedoras (Aranae: Araneidae, Tetragnathidae, Theridiidae) en cafetales del Soconusco, Chiapas, Mexico. Folia Entomologica Mexicana 102:11-20. Jimenez, M.L. 1996. Araneae. Pp. 83-101. In Biodiversidad, Taxonomia y Biogeografia de Artropodos de Mexico: Hacia una Sintesis de su Conocimiento. (J. Llorente, A.N. Garcia-Aldrete & S.E. Gonzalez, eds.). Instituto de Biologia, Universidad Nacional Autonoma de Mexico, Mexico. Kovach, W.L. 1999. MVSP: a multivariate statistical package for Windows, ver. 3.1. Kovach Computing Services, Pentraeth, Wales, UK. 133 pp. Lubin, Y.D. 1978. Seasonal abundance and diversity of web-building spiders in relation to habitat structure on Barro Colorado Island, Panama. Journal of Arachnology 6:31-51. McCaffrey, J.P., M.P. Parrella & R.L. Horsburgh. 1984. Evaluation of the limb-beating method for estimating spider (Araneae) populations on apple trees. Journal of Arachnology 11:363-368. Mooney. H.A., S.H. Bullock & J.R. Ehleringer. 1989. Carbon isotope ratios of plants of a tropical dry forest in Mexico. Functional Ecology 3:137-142. Moreno-Molina, E.B., G.Y. Ibarra Nunez & A. Garcia Ballinas. 2001. Diversidad de aranas en follaje de cacao en el Soconusco, Chiapas, Mexico. Memorias del XXXII Congreso Nacional de la Sociedad Mexicana de Entomologia. Metepec, Puebla, Mexico. 17 pp. Murphy, P.G. & A.L. Lugo. 1986. Ecology of tropical dry forest. Annual Review of Ecology and Systematics 17:67-88. Pinkus-Rendon, M.A., G. Ibarra-Nunez, V. Parra-Tabla, J.A. Garcia-Ballinas & Y. Henaut. 2006. Spider diversity in coffee plantations with different management in Southeast Mexico. Journal of Arachnology 34:104-112. Raizer, J. & M.E.C. Amaral. 2001. Does the structural complexity of aquatic macrophytes explain the diversity of associated spider assemblages? Journal of Arachnology 29:227-237. Riechert, S.E. & C.R. Tracy. 1975. Thermal balance and prey availability: bases for a model relating web-site characteristics to spider reproductive success. Ecology 56:265-285. Robinson, J.V. 1981 . The effect of architectural variation in habitat on a spider community: an experimental field study. Ecology 62:73-80. Rypstra, A.L., P.E. Carter, R.A. Balfour & S.D. Marshall. 1999. Architectural features of agricultural habitats and their impact on the spider inhabitants. Journal of Arachnology 27:371-377. Rzedowski, J. 1 978. La Vegetacion de Mexico. Limusa, Mexico. 432 pp. Samu, F., K.D. Sunderland & C. Szinetar. 1999. Scale-dependent dispersal and distribution patterns of spiders in agricultural systems: a review. Journal of Arachnology 27:325-332. Southwood, T.R.E. 1978. Ecological Methods (2nd edition). Chap- man and Hall, London. 524 pp. Souza, A.L.T. De & R.P. Martins. 2004. Distribution of plant- dwelling spiders: inflorescences versus vegetative branches. Austral Ecology 29:342-349. Souza, A.L.T. De & R.P. Martins. 2005. Foliage density of branches and distribution of plant-dwelling spiders. Biotropica 37:416-420. Turnbull, A.L. 1973. Ecology of the true spiders (Araneomorphae). Annual Review of Entomology 18:305-348. Uetz, G.W., J. Halaj & A.B. Cady. 1999. Guild structure of spiders in major crops. Journal of Arachnology 27:270-280. Vasconcelos-Neto, J., G.Q. Romero & A.J. Santos. 2006. Association of spiders in the Genus Peucetia (Oxyopidae) with plants bearing glandular hairs. Biotropica 39:221-226. Whitmore, C., R. Slotow, T.E. Crouch & A.S. Dippenaar-Schoeman. 2002. Diversity of spiders (Araneae) in a savanna reserve. Northern Province, South Africa. Journal of Arachnology 30:344-356. Wise, E.H. 1993. Spiders in Ecological Webs. Cambridge University Press, Cambridge, UK. 328 pp. Manuscript received 29 January 2005, revised 29 January 2008. 2008. The Journal of Arachnology 36:557-564 Microhabitat preferences for the errant scorpion, Centruroides vittatus (Scorpiones, Buthidae) C. Neal McReynolds: Department of Biology and Chemistry, Texas A&M International University, Laredo, Texas 78041, USA. E-mail: nmcreynolds@tamiu.edu Abstract. Vegetation as a preferred microhabitat for scorpions has rarely been considered despite many Buthidae (the bark scorpions) being non-burrowing errant scorpions that are active on both the ground and vegetation. Microhabitats can serve multiple functions for Centruroides vittatus (Say 1821), but a particular microhabitat can be preferred for a certain function such as a refuge, foraging, or feeding. Observations of microhabitat use by C. vittatus were performed in Laredo, Texas of the Tamaulipan Biotic Province. Comparisons of microhabitat use by C. vittatus at different temperatures or precipitation levels were performed. Foraging and feeding by C. vittatus among microhabitat classes were also compared. The observed use of vegetation by C. vittatus during different seasons was compared to the expected use based on relative abundance of vegetation in the habitat. Air temperature, but not precipitation, had a significant effect on microhabitat use by C. vittatus. Microhabitat had a significant effect on foraging of C. vittatus with caterpillars comprising 34.6% of the prey items and half of the scorpions feeding on caterpillars were in blackbrush ( Acacia rigidula). The lowest proportion of scorpions observed feeding was on the ground (3.8%) and the highest in blackbrush (40.4%). The frequency of C. vittatus among vegetation classes was significantly different compared to the relative abundance of plant species in the plot. Scorpions were observed on prickly pear cactus (Opuntia engelmannii) and strawberry cactus ( Echinocereus enneacanthus) at a higher frequency than expected, and scorpions were observed on guajillo (Acacia berlandieri) and tasajillo (Opuntia leptocaulis) at a lower frequency than expected. The frequency of scorpions on blackbrush was higher than expected during the spring. Vegetation is an important microhabitat for C. vittatus in south Texas. The results indicate the possibility that C. vittatus in south Texas used various plant species to carry prey captured on the ground into vegetation to feed, used blackbrush to forage for caterpillars, and used strawberry and prickly pear cacti as a possible refuge. Keywords: Habitat selection, foraging, refuge, feeding Scorpions utilize a diversity of habitats (Hadley & Williams 1968; Polis 1990). Studies of habitat selection by scorpions have compared soil types for foraging (Polis & McCormick 1986a) or to build burrows (Polis & Farley 1980; Bradley & Brody 1984; Bradley 1988; Smith 1998). The effects of vegetation on scorpions has been considered in association with soil types (Bradley 1988), fire (Smith & Morton 1990), or as refuge from predation or cannibalism (Polis 1980a) including scorpions fleeing under vegetation to avoid preda- tors because of low light levels (Camp & Gaffin 1999). However, vegetation as a preferred microhabitat for scorpions has rarely been considered despite many Buthidae (the bark scorpions) being non-burrowing errant scorpions that are active on both the ground and vegetation (Hadley & Williams 1968; Polis 1990). Centruroides vittatus (Say 1821) (Scorpiones; Buthidae), the striped bark scorpion, has a wide distribution utilizing a number of different habitats (Shelley & Sissom 1995). Studies of habitat use by C. vittatus have already been done in the desert of west Texas (Brown & O’Connell 2000; Brown et al. 2002), in the deciduous forest of Arkansas (Yamashita 2004), and in the chaparral of south Texas (McReynolds 2004). This study will consider if C. vittatus has microhabitat preferences that can increase the fitness of the scorpion in the chaparral of south Texas. Centruroides vittatus can utilize microhabitats for a refuge, foraging, or feeding. Many buthids will use vegetation or debris as a refuge during the day (Polis 1990). Refuges such as burrows can be used to avoid extreme temperatures during the day for many species of scorpions (Hadley 1974), and rocks and cracks in the ground can serve the same function for Centruroides sculpturatus Ewing 1928 (Crawford & Krehoff 1975) and C. vittatus (Brown et al. 2002). Vegetation could also be a refuge from extreme conditions such as high temperatures during the day or low temperatures at night. One possibility is that cacti such as Texas prickly pear cactus (Opuntia engelmannii ), tasajillo (Opuntia leptocaulis) and strawberry cactus (Echinocereus enneacanthus) can be a refuge from these extreme conditions because the high water content in the cacti could provide a buffer from temperature changes due to the high specific heat of water. Centruroides vittatus can climb into vegetation to forage or to feed on prey captured on the ground. Brown & O’Connell (2000) hypothesized that C. vittatus climbs into vegetation because of predator avoidance or higher prey availability. If feeding scorpions carry prey up vegetation to avoid ground predators (such as Lycosidae and Solifugae), then any vegetation can be used as a site to feed assuming that all plant species provides the same protection from ground predators. In addition, feeding scorpions carrying prey up vegetation from the ground should mainly feed on prey that is captured on the ground (Brown & O’Connell 2000). If scorpions are foraging in vegetation, then scorpions can be searching for prey found only in vegetation (e.g., lepidopteran caterpillars) and should prefer vegetation with high availabil- ity of these prey such as blackbrush ( Acacia rigidula) and guajillo ( Acacia berlandieri). Furthermore, scorpions can forage in the vegetation when prey availability in vegetation is higher such as during periods of high precipitation (see Polis 1979, 1980b). 557 558 THE JOURNAL OF ARACHNOLOGY Microhabitats can serve multiple functions for C. vittatus , but a particular microhabitat can be preferred for a certain function. This study will consider how certain conditions can affect microhabitat use by scorpions. Microhabitat use will be compared in relation to temperature and precipitation for possible shifts in activity among microhabitats. Microhabitat use will be compared in relation to prey capture and feeding to determine if scorpions are foraging in vegetation and/or carrying prey from the ground to vegetation. The observed use of trees, shrubs and cacti by C. vittatus during three time periods will be compared to a census of plant species. These comparisons are to determine if microhabitat selection was random or C. vittatus shows a preference for microhabitats during any seasonal periods. METHODS Study animal. — Centruroides vittatus has a wide distribution with Laredo, Texas in the southern portion of the distribution (Shelley & Sissom 1995). Centruroides vittatus is nocturnal with refuges during the day in debris, beneath vegetation, under bark, and in holes in the ground, but C. vittatus and other bark scorpions rarely dig their own burrows (Polis 1990). Scorpions emerge from their refuge only occasionally to forage (Polis 1980b; Bradley 1988; Warburg & Polis 1990). Different sized scorpions can be observed throughout the year with birth of C. vittatus between April and September and age of maturity of 36 to 48 mo (Polis & Sissom 1990). On nights of emergence, C. vittatus is active on the ground and/or in vegetation. Courtship by C. vittatus has rarely been observed and females carrying first instars observed only occasionally in the field (pers. obs.). Voucher specimens of C. vittatus were deposited in the invertebrate collection at Texas A&M International University. Habitat. — This study was done on the campus of Texas A&M International University (27°35'N, 99°26'W), Laredo, Texas. Laredo is in the Tamaulipan Biotic Province that is characterized by low precipitation and high average temper- atures (Blair 1950). The habitat of the research plots can be described as thorny brush (Blair 1950) or chaparral. Vegeta- tion in the plots included blackbrush ( Acacia rigidula ), guajillo ( Acacia berlandieri), honey mesquite (Prosopis glandulosa), Texas prickly pear cactus ( Opuntia engelmannii ), tasajillo ( Opuntia leptocaulis ), strawberry cactus (Echinocereus ennea- canthus ), cenizo (Leucophyllum frutescens), guayacan ( Guaia - cum angustifolium ), leather stem (Jatropha dioica), lotebush ( Ziziphus obtusifolia), Spanish dagger ( Yucca treculeana ), and other species. Three research plots of the campus were studied from 14 September 2000 to 8 August 2002 over 20 nights in 2000, 67 in 2001, 15 in 2002. The study continued from 28 August 2002 to 12 May 2005 in the main research plot over 21 nights in 2002, 52 in 2003, 46 in 2004, and 26 in 2005. At the start, three circular sites of 100 nr were placed in each plot to include different vegetation. Additional sites were placed at random in the main research plot, and these sites were first searched on 1 November 2000. The sites in the other two plots were abandoned 8 August 2002 because of construction nearby and light pollution from streetlights on the campus. Data collection. -Scorpions were observed at night by locating the scorpion fluorescing under ultraviolet light (see Sissom et al. 1990). Observed scorpions were active and either out of or just emerging from their refuges. No data were collected on scorpions in their refuges to avoid destruction of the habitat. Scorpion data were collected after sunset between 19:30 Central Standard Time, USA (CST) at the earliest and 01:00 CST at the latest for an average of two hours per night of observation. Sites were selected at random and searched during a night of observation with a mode of three sites searched per night. Data were collected on all scorpions observed within or near the site. Data collected for each scorpion included date and time of observation, species of scorpion, microhabitat used, if prey was captured or not, and prey taxa. Air temperature was collected each night using a portable weather meter, Kestrel® 3000, from 16 July 2000 to 12 May 2005. Precipitation data were radar estimates for the field site that were provided by the Center for Earth and Environmental Studies, Texas A&M International University from 1 June 2003. Total precipitation for the two weeks just prior to the sample night was used for analysis because precipitation for the prior two weeks showed a significant effect on the prey availability in blackbrush (unpubl. data). All months of a year were sampled, but scorpions were rarely active during December and January. Scorpions can be active during all other months especially when the temperature is above 20° C during the night. Data collection occurred during 94 nights between January-April, 66 nights between May- August, and 115 nights between September-December. The microhabitat data were placed in different classes: ground, grass, blackbrush, guajillo, prickly pear cactus, tasajillo, strawberry cactus, and other vegetation for the comparisons in this paper. If observed on soil, leaf litter, or a rock, the scorpion was considered on the ground. Grasses were not identified to species, but all other plants were identified to species if possible. Other vegetation included small trees that are rarely taller than 2 meters with the exception of a few mesquites and perennial shrubs such as cenizo, guayacan, leather stem, lotebush, and Spanish dagger. Mesquite was included in other legumes instead of other vegetation for the comparison of prey captured in different microhabitats. Annuals were rare in the habitat except for ephemeral wildflowers after heavy rains and scorpions were rarely observed climbing these wildflowers. Prey capture classes Included no prey captured, caterpillars (Lepidoptera larvae), other insects (including adult Lepidoptera), and IGP (intraguild prey including Scorpiones, Araneae, Solifugae, and Chilopoda). Prey capture by scorpions can be observed as scorpions digest externally, thus prey items can be observed in pedipalps or chelicerae (Polis 1979). Census of vegetation. — A census of plant species in the main research plot was performed on four randomly selected sites from the 12 sites in use during the summer of 2001 and the spring of 2002. Each circular site had all trees, shrubs and cacti within the 100 nr area identified to species and counted (Table 1). Grasses and ephemeral wildflowers were not sampled. The proportions of plant species in the four sites can be used to predict the expected frequency of scorpions on vegetation as if there was no preference in vegetation use (Table 1). Only scorpions on live vegetation that were included in the census were included in the comparisons (no scorpions on the ground, grass, ephemeral wildflowers, or dead vegetation). The observed vegetation use of C. vittatus in McREYNOLDS— MICROHABITAT PREFERENCES FOR CENTRUROIDES VITTATUS 559 Table 1. — The number, proportion (%), and estimated density of plant species censused from four random sites in the main research area during the summer of 2001 and spring of 2002. Each site was a circle with an area of 100 nr. Species Common name Number Proportion (%) Density (#/ha) Acacia rigiduia Blackbrush 51 28.3 1275 Acacia berlandieri Guajillo 33 18.3 825 Opuntia leptocaulis Tasajillo 33 18.3 825 Opuntia engelmannii Prickly pear cactus 12 6.7 300 Guaiacum august [folium Guayacan 12 6.7 300 Echinocereus enneacanthus Strawberry cactus 5 2.8 125 Jatropha dioica Leather stem 5 2.8 125 Prosopis glandulosa Honey mesquite 2 1.1 50 Other vegetation 27 15.0 675 Total 180 4500 the main research plot during three time periods was compared to the expected vegetation use. The three time periods were based on a previous analysis of seasonal differences in microhabitat use (McReynolds 2004). Data analyses. — Analysis of contingency tables (Model I) for effects on microhabitat and foraging used the G-test of independence (Sokal & Rohlf 1995). Planned comparisons were performed on a significant association for the contin- gency tables to test predictions on microhabitat preferences. The first planned comparison was among ground and vegetation. The second planned comparison was either to test for differences among vegetation classes that were predicted to be used for foraging ( Acacia spp.), for refuges (cacti) and other vegetation or to test for differences in prey capture among legumes ( Acacia spp. and Prosopis glandulosa) and other vegetation. Other comparisons were performed to complete the orthogonal comparisons. The replicated goodness of fit G- test was used to compare the observed vegetation use by C. vittatus for three time periods to the expected vegetation use based on the census of vegetation (Sokal & Rohlf 1995). RESULTS Effects on microhabitat use. — Air temperature had a significant effect on microhabitat use (Fig. 1). In planned comparisons, ground classes were significantly different from pooled vegetation classes, blackbrush classes were significantly different from guajillo classes, and grass classes were marginally significantly different from other vegetation classes (Fig. 1, Table 2). However, there was no significant difference among Acacia spp., cacti, or other vegetation classes (Fig. 1, Table 2). The proportion of scorpions on vegetation was highest at intermediate temperature class (20-25° C) and lowest at high temperature class (> 30° C) (Fig. 1). Precipi- tation for the two weeks prior to observations had no significant effect on microhabitat use (Fig. 2). Foraging.-- Microhabitat had a significant effect on forag- ing of C. vittatus (Fig. 3). In planned comparisons, ground was significantly different from all vegetation, and legumes were significantly different from other vegetation (Fig. 3, Table 3). Only a very small proportion of scorpions on the ground had prey compared to scorpions in vegetation. The scorpions in the legumes had a high proportion of caterpillars and other insects as prey while the scorpions in other vegetation and cacti had a high proportion of intraguild prey (IGP). The lowest proportion of scorpions observed feeding (/? = 104) was on the ground (3.8%) and the highest in blackbrush (40.4%). Caterpillars were 34.6% of the prey items for C. vittatus , and half of the scorpions observed feeding on caterpillars were in blackbrush. Intraguild prey (IGP) were 17.3% of the prey items for C. vittatus with 9.6% Araneae, 3.8% Scorpiones, 1.9% Solifugae, and 1.9% Chilopoda. Microhabitat preferences. — The proportion of C. vittatus on vegetation was compared to the expected proportion for three time periods (Fig. 4). The expected proportion assumed that scorpions have no preference for vegetation, and the distribution of scorpions on vegetation will be random relative to the abundance of plant species in the research plot (see Table 1 ). The proportion of scorpions on vegetation was significantly different from expected for all three time periods and the pooled data, and the three time periods were significantly heterogeneous (Fig. 4, Table 4). Scorpions were observed on prickly pear and strawberry cacti at a higher frequency than expected for every time period and the pooled data, and scorpions were observed on both guajillo and tasajillo at a lower frequency than expected for every time period and the pooled data. However, the frequency of scorpions on blackbrush was higher than expected during the January-April time period but lower than expected during the May-August time period and only slightly higher than expected during the September-November time class. The heterogeneity between time periods was due to fluctuations in the frequency of scorpions in the blackbrush and other vegetation classes. DISCUSSION Comparisons of scorpions have often noted the lack of activity on vegetation (Bradley 1988; Warburg & Polis 1990). One explanation for this pattern is that adaptation to a specialized habitat (sand) can reduce effectiveness in climbing (see Fet et al. 1998). However, bark scorpions (Buthidae) are known to be active on vegetation (Polis 1990). Microhabitat use of Buthus occitanus (Amoreux 1789) includes juveniles on bushes but not adults '(Skutelsky 1996). In a study by Hadley & Williams (1968), Centruroides sculpturatus pursues prey up and down vegetation and under rocks and is more active than the other scorpion species. Centruroides vittatus in south Texas (present study) uses vegetation at higher frequency than in west Texas (Brown & O’Connell 2000) and in Arkansas (Yamashita 2004). Vegetation is important microhabitat for C. vittatus in Laredo, Texas with 54.1% on trees, shrubs or 560 THE JOURNAL OF ARACHNOLOGY Figure 1 . The number of Centruroides vittatus using different microhabitats among temperature classes. The two Opimtia spp. classes, O. leptocaulis (tasajillo) and O. englemanii (prickly pear cactus), were pooled for the statistical analysis. The frequency of microhabitat use was significantly different among temperature classes (G = 74.79, P < 0.001, df = 18, n = 1252). See Table 2 for planned comparisons among microhabitat classes. cacti; 7.8% on grass and only 38.1% on the ground. In other studies, C. vittatus were observed in trees (9.1%) and grass (10.6%) in Arkansas (Yamashita 2004) and 26.4% climbing vegetation in west Texas (Brown & O’Connell 2000). Scorpion activity and thus microhabitat use can shift because of environmental factors (i.e., temperature and/or precipitation). The environmental factors can have a direct effect on the scorpion activity or indirectly on prey availability (Polis 1980a. 1988 but see Bradley 1988). Microhabitat use of C. vittatus shifted to the ground at high nocturnal tempera- tures. However, there was no support for the prediction that microhabitat use would shift from refuges in cacti to foraging in Acacia spp. with differences in temperature although the use of blackbrush was high during intermediate temperatures relative to guajillio. The high activity of scorpions on the Table 2.- Planned comparisons among microhabitat classes of the contingency table for microhabitat vs. temperature classes. The two Opimtia spp. classes, O. leptocaulis (Tasajillo) and O. englemanii (Prickly pear cactus), were pooled for the statistical analysis. NS = not significant. See Fig. 1. Planned comparisons G df P Ground vs. All vegetation 43.71 3 < 0.001 Cactus vs. Acacia spp. vs. Grass and Other vegetation 8.71 6 NS Opuntia spp. vs. Strawberry cactus 3.77 3 NS Blackbrush vs. Guajillo 11.33 3 < 0.05 Grass vs. Other vegetation 7.27 3 0.1-0.05 Total 74.79 18 < 0.001 ground at high temperatures (> 30° C) during the night does fit the pattern of high activity of C. vittatus on the ground during July and August as previously reported (McReynolds 2004). This can indicate low prey availability in vegetation and relatively higher prey availability on the ground during the hottest period of the year. Microhabitat use of C. vittatus did not shift with precipitation, and there was no evidence that foraging in blackbrush increased with high precipitation as predicted because of the observed increase in caterpillar availability with high precipitation (unpubl. data). One possible reason that activity and foraging behavior does not change with precipitation (and prey availability) is the threat of predation including cannibalism by adults on juveniles (see Polis 1980a, 1980b). Only adult Paruroctonus mesaensis (now Smeringus mesaensis (Stahnke 1957)) have a significant positive correlation with prey availability while the other age classes have a positive but not significant correlation (Polis 1980b). Scorpions can utilize different microhabitats and in particular different vegetation for feeding, foraging or refuge. Scorpion species can feed where the prey was captured, can carry prey to burrow (or other refuge) before feeding, or can carry it into vegetation. For example, Parabuthus pallidus Pocock 1895 carries prey back to the burrow but Parabuthus leiosoma (Ehrenberg 1828) feeds where prey is captured (Rein 2003). Scorpions with prey on vegetation are usually attributed to scorpions carrying prey from the ground into vegetation to feed (Polis 1979; Brown & O’Connell 2000). Only 3.8% of the scorpions with prey were on the ground in south Texas but many prey of C. vittatus were usually McREYNOLDS— MICROHABITAT PREFERENCES FOR CENTRUROIDES VITTATUS 561 220 -r 200- 180 160 140 I120 O 100 5 Precipitation (cm) Figure 2. — The number of Centruroid.es vittatus using different microhabitats among classes of total precipitation for the prior two weeks. The frequency of microhabitat use was not significantly different among precipitation classes (G = 7.37, not significant, df — 14, n = 985). Figure 3. — The number of Centruroides vittatus using different microhabitats among prey capture classes. The frequency of microhabitat use was significantly different among prey capture classes (G = 112.17, P < 0.001, df = 12, n = 1890). See Table 3 for planned comparisons among microhabitat classes. 562 THE JOURNAL OF ARACHNOLOGY Table 3. — Planned comparisons among microhabitat classes of the contingency table for microhabitat vs. prey capture classes. NS = not significant. See Fig. 3. Planned comparisons G df P Ground vs. All vegetation 83.77 3 < 0.001 Legumes vs. Cactus and Other vegetation 14.92 3 < 0.01 Blackbrush vs. Other legumes 6.53 3 NS Cactus vs. Other vegetation 6.95 3 NS Total 112.17 12 < 0.001 observed on the ground (e.g., many of the intraguild prey) (pers. obs.). Intraguild prey (including cannibalism) were 17.3% of the prey in south Texas as compared to 27.91% of spider prey and 9.30% cannibalism for C. vittatus in Arkansas (Yamashita 2004). These data indicate that C. vittatus in south Texas do carry prey captured on the ground into vegetation to feed. One possible function of moving prey (including intraguild prey) to vegetation is to avoid intraguild predation (see Bradley & Brody 1984; Polis & McCormick 1986b, 1987). Another possible function can be to avoid scavenging ants. Ants were observed causing feeding scorpions to move (pers. obs. by E. Lopez and C.N. McReynolds). Scorpions can also move between vegetation. A scorpion was observed feeding on a caterpillar on tasajillo but near a blackbrush where the caterpillars are available (pers. obs.). Perhaps this is to avoid ants or predators (e.g., other scorpions). Foraging in vegetation has not been considered important for scorpions (Polis 1979) except for errant scorpions (McCormick & Polis 1990) and by juveniles (e.g., juvenile Buthus occitanus ambush prey in vegetation but not adults (Skutelsky 1996)). There was no evidence of foraging in vegetation by C. vittatus in west Texas (Brown & O’Connell 2000). However, the diet of C. vittatus in Laredo, Texas includes a number of items that have a high availability in trees or shrubs. Caterpillars were an important prey item for the scorpions in south Texas and most of the caterpillar taxa captured by C. vittatus were available in blackbrush (unpubl. data). Caterpillars are rarely reported as an important item in the diet for scorpions (see McCormick & Polis 1990). Caterpillars were only 1.4% of the diet for P. mesaensis (Polis 1979), and no caterpillars were reported for C. vittatus in west Texas (Brown & O’Connell 2000) but 1 1 .6% of the prey for C. vittatus in Arkansas (Yamashita 2004) and 34.6% in south Texas (present study). It is predicted that scorpions foraging in blackbrush (and other legumes) will increase when caterpillar availability increases. However, scorpions prefer blackbrush only in January-April and there is no shift to blackbrush with high precipitation. If caterpillar availability is higher with the blooming of blackbrush in March and April, then this can explain why there is not an overall preference for blackbrush but there is a higher proportion of prey captured during March and April (McReynolds 2004) and the proportion of scorpions on blackbrush is higher than expected during the January-April time period. January-April May-August September-November Pooled Expected Figure 4. The proportion (%) of Centruroides vittatus on vegetation during each seasonal time period compared to the expected proportion (%) on vegetation. The observed frequency of scorpions on vegetation from January to April (n = 174), from May to August (/; = 166), from September to November (/? = 415) and for the pooled data {n = 755) were compared to the expected frequency based on relative abundance of vegetation (see Table I) in the replicated goodness of fit test (see Table 4). McREYNOLDS— MICROHABITAT PREFERENCES FOR CENTRUROIDES VITTATUS 563 Table 4. — Replicated goodness of fit test for microhabitat classes comparing the observed frequency of scorpions on vegetation to the expected frequency based on relative abundance of vegetation (see Table 1) during three seasonal time periods. See Fig. 4. Time Periods G df P January to April 113.74 5 < 0.001 May to August 130.43 5 < 0.001 September to November 293.61 5 < 0.001 Pooled 473.47 5 < 0.001 Heterogeneity 64.31 10 < 0.001 Total 537.77 15 < 0.001 Most scorpion species dig a burrow, but many buthids do not dig their own burrow but use holes, space beneath rocks, and openings under bark and below vegetation as diurnal refuges (Polis 1990). Centruroides vittatus in west Texas have patchy distribution under rocks as diurnal refuges (Brown et al. 2002). Scorpions have been observed entering cracks or holes in the ground in south Texas (pers. obs.), but rocks are not available in the main research plot of this study. An alternative refuge for C. vittatus in south Texas can be a cactus because the high water content can provide a buffer from temperature changes due to the high specific heat of water. Scorpions have been observed going under the pads of a prickly pear cactus or down the openings between the stems of a strawberry cactus (pers. obs.). This can explain the higher than expected frequency of C. vittatus on strawberry cactus and prickly pear cactus in the research plot despite the low frequency of scorpions feeding on cacti and the low probability of foraging success because of the low prey availability on cacti (pers. obs.). This is assuming that C. vittatus is a central place forager (Orians & Pearson 1979) and there is a high probability that the scorpion would be near its refuge after emerging for the night (see Polis et al. 1985). However, C. vittatus has a lower than expected frequency in tasajillo. Perhaps tasajillo with thin stems and the more treelike structure does not provide the refuge that the prickly pear cactus or the strawberry cactus can provide. Sampling of potential refuges during the day at different temperatures will be required to establish diurnal refuge preferences of C. vittatus in south Texas (see Brown et al. 2002). The high frequency of scorpions with prey in vegetation and low frequency on the ground indicates that C. vittatus carry prey caught on the ground into vegetation to feed. Foraging by C. vittatus especially in blackbrush for caterpillars has been demonstrated. However, the prediction that foraging scorpi- ons will show a preference for blackbrush and other legumes was not supported. The frequency of scorpions in blackbrush did not increase with higher precipitation as is predicted because high precipitation increases availability of caterpillars in blackbrush. The frequency of scorpions in blackbrush was higher than expected in comparison to relative abundance of plant species only in January through April. It is possible that prey availability is higher early in the year especially March and April, but there is no evidence yet supporting this prediction. The results show that C. vittatus in south Texas used strawberry and prickly pear cacti with a higher than expected frequency in all time periods. However, the low frequency of scorpions with prey on cacti suggests that this preference for cacti was not for feeding or foraging, and perhaps scorpions were utilizing the cacti for another function such as a diurnal refuge. ACKNOWLEDGMENTS I thank my undergraduate research students from Fail 2000 to Spring 2005 for assistance in the field to collect the data for the paper. They included Elaine Boteilo, Ezequiel Chapa, Yareth Coutino, Ruperto Contreras III, Luis Ricardo de Leon, Paula Flores, Nicolas Gallegos, Jose Hernandez, Juvenal Herrera IV, Mike Herrera HI, Eduardo Lopez, Ruben Mendez, Erica Morales, Rocio Moya, Metzli Ortega, Julianna Quintanilla, Oscar Ramos, Amede Rubio, and Israel Salinas. In particular, I thank Metzli Ortega for the census of plant species and Eduardo Lopez for his observations on the feeding behavior of scorpions and their interactions with ants. Thanks to Gene McReynolds for his modifications of the ultraviolet lights and to Michael Daniel for a discussion on the potential of cactus as a refuge. I also thank Dan Mott, Julianna Quintanilla, Oscar Ramos, Soren Toft, Tsunemi Yamashita, Fernando Quintana, and two anonymous reviewers for suggestions to improve the paper immeasurably. Financial support was provided by grants from Texas A&M Interna- tional University. The scorpion research at Texas A&M International University was started at the suggestion of the late Gary Polis. LITERATURE CITED Blair, W.F. 1950. The biotic provinces of Texas. Texas Journal of Science 2:93-1 17. Bradley, R.A. 1988. The influence of weather and biotic factors on the behavior of the scorpion (. Paruroctonus utahensis). Journal of Animal Ecology 57:533-551. Bradley, R.A. & A.J. Brody. 1984. Relative abundance of three vaejovid scorpions across a habitat gradient. Journal of Arachnol- ogy 11:437-440. Brown, C.A., J.M. Davis, D.J. O’Connell & D.R. Formanowiz, Jr. 2002. Surface density and nocturnal activity in a west Texas assemblage of scorpions. Southwestern Naturalist 47:409-419. Brown, C.A. & D.J. O’Connell. 2000. Plant climbing behavior in the scorpion Centruroides vittatus. American Midland Naturalist 144:406-418. Camp, E.A. & D.D. Gaffrn. 1999. Escape behavior mediated by negative photoaxis in the scorpion Paruroctonus utahensis. Journal of Arachnology 27:679-684. Crawford, C.S. & R.C. Krehoff. 1975. Die! activity in sympatric populations of the scorpions Centruroides scidpturatus (Buthidae) and Diplocentrus spitzeri (Diplocentridae). journal of Arachnology 2:195-204. Fet, V., G.A. Polis & W.D. Sissom. 1998. Life in sandy deserts: the scorpion model. Journal of Arid Environments 39:609-622. Hadley, N.F. 1974. Adaptationai biology of desert scorpions. Journal of Arachnology 2:11-23. Hadley, N.F. & S.C. Williams. 1968. Surface activities of some North American scorpions in relation to feeding. Ecology 49:726-734. McCormick, S.J. & G.A. Polis. 1990. Prey, predators, and parasites. Pp. 294-320. In The Biology of Scorpions. (G.A. Polis, ed.). Stanford University Press, Stanford, California. McReynolds, C.N. 2004. Temporal patterns in microhabitat use for the scorpion Centruroides vittatus (Scorpiones, Buthidae). Euscor- pius 17:35-45. Orians, G.H. & N.E. Pearson. 1979. On the theory of central place foraging. Pp. 155-177. In Analysis of Ecological Systems. (D.J. 564 THE JOURNAL OF ARACHNOLOGY Horn, R. Mitchell & G.R. Stair, eds.). Ohio State University Press, Columbus, Ohio. Polis, G.A. 1979. Prey and feeding phenology of the desert sand scorpion Paruroctonus mesaensis (Scorpionidae: Vaejovidae). Journal of Zoology (London) 188:333-346. Polis, G.A. 1980a. The effect of cannibalism on the demography and activity of a natural population of desert scorpions. Behavioral Ecology and Sociobiology 7:23-35. Polis, G.A. 1980b. Seasonal patterns and age-specific variation in the surface activity of a population of desert scorpions in relation to environmental factors. Journal of Animal Ecology 49:1-18. Polis, G.A. 1988. Foraging and evolutionary responses of desert scorpions to harsh environmental periods of food stress. Journal of Arid Environments 14:123-134. Polis, G.A. 1990. Ecology. Pp. 247-293. In The Biology of Scorpions. (G.A. Polis, ed.). Stanford University Press, Stanford, California. Polis, G.A. & R.D. Farley. 1980. Population biology of desert scorpion: survivorship, microhabitat, and the evolution of life history strategy. Ecology 61:620-629. Polis, G.A. & S.J. McCormick. 1986a. Patterns of resource use and age structure among species of desert scorpion. Journal of Animal Ecology 55:59-73. Polis, G.A. & S.J. McCormick. 1986b. Scorpions, spiders and solpugids: predation and competition among distantly related taxa. Oecologia 71:111-116. Polis, G.A. & S.J. McCormick. 1987. Intraguild predation and competition among desert scorpions. Ecology 68:332-343. Polis, G.A., C.N. McReynolds & R.G. Ford. 1985. Home range geometry of the desert scorpion Paruroctonus mesaensis. Oecologia 67:273-277. Polis, G.A. & W.D. Sissom. 1990. Life history. Pp. 161-223. In The Biology of Scorpions. (G.A. Polis, ed.). Stanford University Press, Stanford, California. Rein, J.O. 2003. Prey capture behavior in the East African scorpions Parabuthus leiosoma (Ehrenberg, 1828) and P. pallidus Pocock, 1895 (Scorpiones: Buthidae). Euscorpius 6:1-8. Shelley, R.M. & W.D. Sissom. 1995. Distributions of the scorpions Centruroides vittatus (Say) and Centruroides hentzi (Banks) in the United States and Mexico (Scorpiones, Buthidae). Journal of Arachnology 23:100-1 10. Sissom, W.D., G.A. Polis & D.D. Watt. 1990. Field and laboratory methods. Pp. 445-461. In The Biology of Scorpions. (G.A. Polis, ed.). Stanford University Press, Stanford, California. Skutelsky, O. 1996. Predation risk and state-dependent foraging in scorpions: effects of moonlight on foraging in the scorpion Buthus occitanus. Animal Behaviour 52:49-57. Smith, G.T. 1998. Density of the burrowing scorpion Urodacus armatus (Scorpiones; Scorpionidae) in relation to vegetation types: implications for population decline following agricultural clearing. Pacific Conservation Biology 4:209-214. Smith, G.T. & S.R. Morton. 1990. Responses by scorpions to fire- initiated succession in arid Australian Spinifex grasslands. Journal of Arachnology 18:241-244. Sokal, R.R. & F.J. Rohlf. 1995. Biometry, 3rd edition. W.A. Freeman and Company, New York. 887 pp. Warburg, M.R. & G.A. Polis. 1990. Behavioral responses, rhy- thms, and activity patterns. Pp. 224—246. In The Biology of Scorpions. (G.A. Polis, ed.). Stanford University Press, Stanford, California. Yamashita, T. 2004. Surface activity, biomass, and phenology of the striped scorpion, Centruroides vittatus (Buthidae) in Arkansas, USA. Euscorpius 17:25-33. Manuscript received 29 January 2007, revised 20 January 2008. 2008. The Journal of Arachnology 36:565-573 Observations on phenology and overwintering of spiders associated with apple and pear orchards in south-central Washington Eugene R. Miiiezky, David R. Horton and Carrol O, Calkins: Yakima Agricultural Research Laboratory, USDA-ARS, 5230 Konnowac Pass Road, Wapato, Washington 98951, USA. E-mail: Gene.Msliczky@ars.usda.gov Abstract. Beating tray and sweep net samples from apple and pear orchards in south-central Washington State were used to obtain information on life history and phenology of orchard-associated spiders. Cardboard shelters placed in the orchards in the fall and collected during the winter yielded information on spider overwintering. Data were obtained for 43 species in 28 genera and 12 families. The eight most abundant species were Pelegrina aeneola (Curtis 1892), Meioneta filhnorana (Chamberlin 1919), Oxyopes scalaris Hentz 1845, Theridion neomexicanum Banks 1901, Misumenops lepidus (Thorell 1877), Xysticus cunctator Thorell 1877, Philodromus cespitum (Walckenaer 1802), and Sassacus papenhoei Peck ham & Peckham 1895. Each was represented by more than 690 specimens. Salticidae, Philodromidae, and Linyphiidae were represented by the largest number of species. Most species appear to have univoltine life cycles in the study area. Species matured at different times during the season between spring and fall. Twenty-seven species utilized cardboard shelters for overwintering, but some common spiders failed to do so and apparently use alternative locations. Some species overwintered in a broad range of developmental stages, whereas other species overwintered in only one or two instars. Keywords: Araneae, life history, seasonality Spiders have been of interest as predators of arthropod pests in orchards for at least 50 years. Chant (1956) found differences between the spider faunas of insecticide treated compared to untreated orchards and listed the natural (i.e., non-orchard) habitats in which many species occurred. Don- dale (1956, 1958) recorded 77 species from apple orchards in Nova Scotia, Canada, provided data on abundance, and observed spiders feeding on apple pests. Similar studies were conducted over the next 35 years: Dondale et al. (1979) and Bostanian et al. (1984) in Canada; Specht & Dondale (1960), Legner & Oatman (1964), and McCaffrey & Horsburgh (1980) in the United States; Mansour et ah (1980a) in Israel; Dondale (1966) in Australia; and Hukusima (1961) in Japan. The efficacy of limb tapping for estimating spider populations in apple orchards was investigated by McCaffrey et al. (1984). Interest in the spider fauna of apple orchards has continued in recent years as pest management programs that rely less heavily on broad-spectrum insecticides are employed with greater frequency, improving the prospects for significant impact by natural enemies on pest control. Studies have examined the tree canopy fauna (Olszak et al. 1992; Wisniewska & Prokopy 1997; Brown et ah 2003; Pekar & Kocourek 2004), the soil surface fauna (Bogya & Marko 1999; Pekar 1999a; Epstein et al. 2000), and the fauna of the herbaceous layer (Pekar 1999b; Bogya et al. 2000; Miiiezky et al. 2000). Overwintering spiders (Bogya et ah 2000; Pekar 1999c; Horton et al. 2001) and the bark dwelling fauna (Bogya et ah 1999b) have also received attention. Regional geographic effects were found to be most important in determining the composition of spider assemblages in orchards in different areas (Bogya et ah 1999a). The phenologies of numerous spider taxa have been documented by field and laboratory studies (Merrett 1967, 1968; Aiken & Coyle 2000; Stiles & Coyle 2001). Annual, biennial, and intermediate length life cycles have been described (Toft 1976, 1978; Wise 1984), and in some cases a species’ life cycle has been found to vary within its range. probably influenced by local climatic conditions (Dondale 1961; Putman 1967). Schaefer (1977) distinguished five types of life cycle among spiders from a north temperate region, studied their adaptive strategies for surviving the winter, and noted the importance of the cold season in synchronizing a species’ life cycle. Here we report on spiders that are found in apple and pear orchards in Yakima County, Washington, an important fruit growing area of the state. Life cycles of several common, orchard-associated species are described, and seasonal distri- bution data for adults and penultimate stage males of less common species are presented. Information on species that utilized artificial shelters for overwintering is also given. Information of this kind may be useful in assessing the potential contribution of spiders to orchard pest control and in scheduling pesticide applications to minimize adverse impact on spiders. METHODS Study orchards. — Sampling was conducted in 42 orchards (apple and pear), all of which were located in Yakima County, Washington. All orchards were within a radius of 46 km of the city of Yakima. Some study orchards have been removed from production since completion of the study. The following are latitudes and longitudes for 8 orchards located at the periphery of the study area: 46.4529°N, 120.2292°W; 46.5029°N, 1 20.1 667° W; 46.5835°N, 120.3501°W; 46.5185°N, 1 20.4235 AV; 46.7432°N, 120.7738°W; 46.6618°N, 120.7557°W; 46.47 15°N, 120.3834 W; 46.3106°N, 120.1236°W. All other orchards fell within the area demarcated by the peripheral orchards. Insect pest management in the study orchards ranged from conventional programs based on synthetic, broad-spectrum insecticides (e.g., azinphos-methyl) to state-certified organic programs in which use of synthetic insecticides is prohibited. The codling moth Cydia pomonella (Linnaeus 1758) is the key pest of apple and pear in this region. The study was conducted from 1996 to 2001, but not all orchards were sampled each year. Nine apple orchards were 565 566 THE JOURNAL OF ARACHNOLOGY sampled in 1996. The following year three pear orchards were sampled in addition to the nine apple blocks. The maximum number of orchards sampled in one year was 19 in 1999. Orchard size was 0.5 to 32 ha. Sampling. — Arboreal spiders were sampled with a beating tray, 0.45 nr in area (Bioquip Products, Gardena, CA). One limb 1-2 m above ground on each of 25 trees (15 trees in 1999) was struck three times with a heavy rubber hose to dislodge spiders. Most specimens were promptly preserved in 70% isopropyl alcohol, but selected specimens were saved and reared (see below). Trees in all parts of an orchard were sampled while walking a winding path. The sampling period usually included April to October. Samples were collected every 1-2 weeks in 1996-1998 and 2000 and monthly in 1999 and 2001. All specimens taken during this study are held at the Yakima Agricultural Research Laboratory (USDA-ARS) in Wapato, Washington, USA. Sweep net sampling (net diameter = 38 cm) of the understory vegetation was done in 1996 and 1997 in the same orchards that were monitored with beat trays. An 180° swing of the net constituted a sweep, and 25 sweeps per sample were taken while walking a winding path so as to sample in all parts of the orchard. The sampling periods were late June to late October 1996 and mid-May to mid-October 1997. Samples were taken every 1-2 weeks with longer intervals after an orchard was mowed. Spiders were also collected by hand on an irregular basis when chanced upon or during an occasional more serious search. Generally, one or a few specimens of interest were collected. Immatures were reared to obtain a positive identification if necessary. These data were used to supplement beat tray and sweep net data. Cardboard shelters of two types were used to collect overwintering spiders. The first consisted of a bundle of ten, 12.5 cm X 17.5 cm sheets of cardboard (flute size ~ 4 mm X 5 mm) tied to the lowest crotch of a tree. The second was a 7.6 cm wide strip of cardboard wrapped once around the trunk of a tree 0. 5-1.0 m above ground. Shelters were set out in September and October, retrieved in December or January, and stored in a cold room until processed. The number of shelters set out and the number of orchards sampled varied from year to year. Incidental to a 1998 study of apple bin use by overwintering codling moth larvae, spiders that had also overwintered in the bins were collected. Bins were removed from the field after harvest, held in a cold room at 0.6°-1.7° C until late January, and then placed in a greenhouse at 2 1 °-24° C. Spiders were collected as they emerged from the bins. Sample processing. —Selected specimens were reared for positive identification. They were held in 35 ml plastic cups and provided water and prey of appropriate size weekly. Field captured Lygus sp. (Hemiptera: Miridae) and laboratory reared Drosophila sp. (Diptera: Drosophilidae) were readily consumed by most species. Once a familiarity with the local fauna was acquired it was possible to identify a majority of immatures to species. Specimens of definite developmental stage (e.g., penultimate female, antepenultimate male, etc.) mentioned in Results in all cases refer to reared individuals. Immatures were sorted into small, medium, and large size classes on a species-by-species basis. Since many spiders pass through five to seven nymphal instars, this roughly corre- sponds to instars one, two, and three for small, four and five for medium, and six and seven for large immatures. Although somewhat arbitrary, this allowed an estimate of the age distribution of immatures through the season. Penultimate males were readily distinguished by their enlarged pedipalps. Theridion , Erigone, and Meioneta, because of their small size, were sorted into a single class of immatures, in addition to penultimate males and adults; antepenultimate male Erigone and Meioneta were distinguishable based on a slight enlarge- ment of the pedipalps. Data presentation. — Data from beat tray and sweep net collections were combined to obtain monthly totals. These data are presented graphically for eight abundant species to show the proportion of different developmental stages in each month’s collection. Note that a small number of beat tray samples (17) were taken during the first week of November. These data were pooled with the extensive October data for the graphs. Few specimens were taken in November because leaf fall was well underway and spiders had begun a general movement out of the trees. Information for less abundant species is summarized in the Tables which also include data from hand collections when available. Overwintering data from all years were combined. RESULTS Salticidae. — Adults of both sexes of Pelegrina aeneola (Curtis 1892) were abundant in April, May, and June (Fig. 1). Thereafter males were rare although females were present through October. Females with egg sacs were most common during May and June, but an egg sac containing undispersed first instar nymphs was collected as late as 8 September 2000. A female collected with her egg sac on 3 June 1997 produced a second clutch of eggs in the laboratory. Small immatures were present in all months but comprised over 85% of the population in July. Penultimate males and large nymphs, which included penultimate females, were most abundant in October. Large immatures and penultimate males were the principal overwintering stages although small and medium-sized immatures were well represented (Table 1). Penultimate males were uncommon in the trees during April, but five adults hand-collected in the litter in early April 1997 may indicate that many penultimate males undergo their final molt in this location and then move up into the trees. Pelegrina aeneola appears to have an annual life cycle in the study area. The phenology of Sassacus papenhoei Peckham & Peckham 1895 was similar to that of P. aeneola but lagged about a month behind (Fig. 1). Adults were most abundant during June, and small immatures dominated the population during August. The few individuals found in overwintering refuges were small and medium-sized immatures (Table 1) as were the few specimens taken in beating tray samples during April. An annual life cycle in the study area is indicated for S. papenhoei. Four species of large jumping spiders in the genus Phidippus occurred in the orchards. Collection data, summarized in Table 2, indicates annual life cycles for all four but with different maturation times. The first two or three nymphal instars were very similar in appearance and could not reliably be sorted to species. Phidippus comatus Peckham & Peckham 1901 matured during the summer. A female guarding her egg MILICZKY ET AL.— ORCHARD SPIDERS 567 Pelegrina aeneola Xysticus cunctator 100% 204 210 1341 1646 926 1036 Sassacus papenhoei 37 279 199 105 Misumenops lepidus 14 553 340 135 m ■w; Oxyopes scalaris Philodromus cespitum F ESM ^sM □ Lg « Med QSm F SM ^sM QLg SMed OSm Figure I. - -Percentage of adults and different sized immatures of six common spider species found by month in combined beating tray and sweep net collections. F = female; M = male; sM = penultimate male; Lg = large immature; Med = medium sized immature; Sm = small immature. Number at the top of the column is the total number of specimens taken for the month. sac was found among a group of ripening pears on 19 August 1998, and a female with a hatched egg mass was found in a codling moth pheromone trap on 29 October 1996. Three small immatures collected on 26 September 2003 yielded a recognizable P. comatus after two molts, a male after five molts, and a penultimate female after six molts. Phidippus clarus Keyserling 1885 also appears to mature during the summer (Table 2). Phidippus audax (Hentz 1845) appears to mature in the spring. Specimens from overwintering shelters (Table 1) were mostly medium and large immatures and. of the 42 individuals that overwintered in apple bins, 26 were penultimate females and 12 were penultimate males. Phidippus johnsoni (Peckham & Peckham 1883) also appears to mature in the spring. Overwintering and seasonal collection data for other Salticidae is summarized in Tables 1 and 2 respectively. Oxyopidae. — Oxyopes scalaris Hentz 1845 was infrequently collected from March to June, but its numbers increased substantially in July and August. Medium-sized immatures made up 50% of the population in September and 73% in October, but large immatures were present in only small numbers both months. Miscellaneous collections included 2 females, 2 penultimate females, and 1 antepenultimate female on 26 May 2000; 1 female and 1 penultimate female on 30 May 2000; and 1 penultimate male on 1 June 2000. Oxyopes scalaris appears to mature in the spring and early summer and to be univoltine in the study area. Thomisidae. — Immature Xysticus cunctator Thorell 1877 were commonly taken in the beat tray samples. Adults, however, were rare in the trees (one female out of 609 specimens) but were more frequently swept from the understory vegetation (five males and four females out of 288 specimens). Adults were most numerous in May according to sweep net and beat tray collections, but hand collected adults of both sexes were taken in April. New generation 568 THE JOURNAL OF ARACHNOLOGY Table 1. — Summary of spider overwintering stages: sm., med., Ig. = small, medium, and large immatures (large immatures included antepenultimates of both sexes and penultimate females); sM = penultimate males; M = adult male; F = adult female. Additional observations: one med. P. cespitum molted three times to an adult female; the sm. P. insperatus molted four times to a male and seven of the med.’s molted three times each yielding five males and two females; one sM P. alascensis was reared to the adult stage. Spider species sm. Overwintering stage med. Ig. sM M F Salticidae Pelegrina aeneola (Curtis 1892) 39 87 227 192 1 1 Sassacus papenhoei Peckham & Peckham 1 895 5 5 Sassacus vitis (Cockerell 1894) 4 3 1 Phidippus audax (Hentz 1845) 1 10 12 4 Phanias watonus (Chamberlin & Ivie 1941) 21 49 23 4 10 28 Salticus scenicus (Clerck 1757) 5 24 34 46 1 Oxyopidae Oxyopes scalaris Hentz 1 845 2 Thomisidae Xysticus cunctator Thorell 1877 1 Misumenops lepidus (Thorell 1877) 29 48 5 34 Philodromidae Philodromus cespitum (Walckenaer 1802) 310 83 10 1 Philodromus insperatus Schick 1965 1 9 Philodromus californicus Keyserling 1884 9 129 1 Philodromus rufus Walckenaer 1826 25 269 7 Philodromus speciosus Gertsch 1 934 2 10 1 3 5 Philodromus alascensis Keyserling 1884 1 2 Tibellus oblongus (Walckenaer 1802) 27 16 4 Ebo pepinensis Gertsch 1933 3 3 2 Clubionidae Cheiracanthium mildei L. Koch 1864 104 210 37 40 Cheiracanthium inclusum Hentz 1847 1 1 Corinnfdae Phrurotimpus borealis Emerton 1911 14 203 140 Anyphaenidae Anyphaena pacifica (Banks 1896) 2 6 6 Theridlidae Steatoda hespera Chamberlin & Ivie 1933 15 31 4 1 3 2 Linyphiidae Meioneta fillmorana (Chamberlin 1919) 1 Erigone spp. 1 Pityohyphantes minidoka Chamberlin & Ivie 1943 11 19 3 1 1 Tetragnathidae Tetragnatha laboriosa Hentz 1850 6 1 1 Dictynidae Dictyna coloradensis Chamberlin 1919 35 21 1 spiderlings appeared in May and remained abundant through a few weeks later than in X. cunctator . Penultimate males first August (Fig. 1). Antepenultimate males, as indicated by appeared in August and were abundant in September, rearing, were present as early as July. Penultimate females October, and the following April Penultimate males and (reared individuals) and penultimate males (Fig. 1) were present by September. Larger immatures, including penulti- mates of both sexes, would therefore be the primary overwintering stages. This species appears to have an annual life cycle in the study area. The phenology of Misumenops lepidus (Thorell 1877) was similar to that of X. cunctator although new generation spiderlings did not appear in large numbers until July (Fig. 1), other immature stages overwintered (Table 1). Misumenops lepidus appeared to have an annual life cycle. Philodromidae. — The phenology of Philodromus cespitum (Walckenaer 1802) (Fig. 1) was similar to that of X. cunctator and M. lepidus but was shifted later into the season. Adult females were most common in June and July, and small spiderlings, which could be found in all months, were most abundant from August to October. Five females with egg sacs MILICZKY ET AL.— ORCHARD SPIDERS 569 Table 2. — Summary by month of beating tray, sweep net, and hand collections of less common spider species: numbers of penultimate males, males, and females, respectively, are given. Additional observations: Penultimate female E. militaris were collected in April. May, September, and October; two antepenultimate female X. gulosus collected in July, one female collected in December; one penultimate female P. insperatus collected in June, one female collected in June laid three egg clutches in the lab; one antepenultimate female P. californicus collected in April, one antepenultimate female and 26 other large immatures collected in October; three antepenultimate male A. trifasciata collected in July and one in August, 2 penultimate females collected in August. Spider species April May June July Aug Sept Oct Salticidae Eris militaris (Hentz 1845) 0-2-0 2-2-11 0-0-4 1-0-4 3-0-1 0-1-0 1-3-0 Phanias watonus (Chamberlin & Ivie 1941) 0-0-2 Phidippus audax (Hentz 1845) 0-1-0 0-0-1 0-1-1 0-0-1 10-0-1 Phidippus darns Keyserling 1885 1-7-11 0-1-1 0-0-3 0-0-1 Phidippus comat us Peckham & Peckham 1901 1-0-0 9-1-1 1-11-8 0-4-4 0-1-0 0-0-1 Phidippus johnsoni Peckham & Peckham 1883 Sasscicus vitis (Cockerell 1894) Salticus scenicus (Clerck 1757) 0-3-2 1-0-0 0-1-1 1-0-0 0-2-0 1-0-0 2-1-0 Thomisidae Xysticus gulosus Keyserling 1880 1-0-0 0-0-1 0-1-3 Philodromidae Philodromus insperatus Schick 1965 Philodromus californicus Keyserling 1884 2-0-0 0-2-2 0-0-4 3-0-0 Philodromus rufus Walckenaer 1826 0-2-0 Tibellus oblongus (Walckenaer 1802) 0-1-2 0-0-1 1-9-9 1-1-7 0-0-1 Tibellus asiaticus Kulczyn’ski 1908 0-0-1 Clubionidae Cheiraccmthium inclusion Hentz 1847 2-1-0 0-0-1 Corinnidae Castianeira longipalpa Hentz 1847 1-0-0 Anyphaenidae Anyphaena pacifica (Banks 1896) Linvphiidae 0-0-1 0-0-1 3-0-0 Spirembolus mundus Chamberlin & Ivie 1933 7-0-0 14-0-0 1 1-0-0 1-1-0 0-1-0 Tenuiphantes tenuis (Black wall 1852) 0-0-1 0-2-2 0-0-1 0-0-2 0-1-1 0-1-0 Collinsia ksenius (Crosby & Bishop 1928) 0-0-1 0-5-3 0-3-1 0-1-3 Walckenaeria subspiralis Millidge 1983 Pityohyphantes minicloka Chamberlin & Ivie 1943 0-1-0 0-5-0 0-1-0 0-1-0 0-2-4 Tetragnathidae Tetragnatha laboriosa Hentz 1850 5-4-2 0-2-2 4-14-11 1-0-3 1-0-2 Tetragnatha versicolor Walckenaer 1841 Araneidae 0-2-0 0-1-1 Argiope trifasciata (Forsskal 1775) 1-0-0 3-1-0 0-0-2 were collected in July, and a female with an egg sac containing summarized in Table 3. All sizes of C. mildei immatures undispersed spiderlings was found in September. Small and medium sized immatures most commonly overwintered (Table 1), and an annual life cycle in the study area is indicated. Tibellus oblongus (Walckenaer 1802) occurred in many orchards. While primarily an inhabitant of the understory vegetation (212 of 256 specimens), individuals were occasion- ally found in the trees. It overwintered as immatures of various sizes (Table 1) and is probably univoltine in the study area. Overwintering and seasonal collection data for less common species in the Philodromidae are given in Tables 1 and 2 respectively. Clubionidae. — Cheiraccmthium mildei L. Koch 1864 oc- curred in many study orchards and appears to have an annual life cycle in the study area. Collection data for the species are overwintered including penultimate males and females (Ta- ble 1). Cheiraccmthium inclusion Hentz 1847 was collected in only one of our study orchards but appears to have a phenology similar to that of C. mildei. Small immatures were most abundant in July (15 specimens) and August (20), medium-sized immatures were most abundant in September (19) and October (17), and large immatures were most abundant in October (13). Theridiidae. — Theridion neomexicanum Banks 1901 was a common, primarily arboreal species that is clearly univoltine in the study area (Fig. 2). Males were first noted in May, but both sexes were most abundant in June and July. A male and a penultimate female were found together in a web on an apple leaf on 1 1 June 1998, females with egg sacs were collected on apple leaves on 10 July 1998 and 27 July 1999, and a female 570 THE JOURNAL OF ARACHNOLOGY Table 3. — Seasonal occurrence of Cheiracanthium mildei based on combined beat tray, sweep net, and hand collections. Spider stages: small, medium, large = small, medium, and large immatures (large immatures included antepenultimate nymphs of both sexes and penultimate females); sM = penultimate males. Additional observa- tions: two of the July and one of the September females were guarding egg sacs. Month of collection Developmental stage small medium large sM Male Female May 1 1 1 June 1 1 1 July 28 5 3 1 3 August 29 28 2 September 18 29 14 5 3 October 19 44 23 10 with a vacated egg sac was found on 6 August 1999. Penultimate males were not observed by October, and overwintering must therefore occur as immatures smaller than penultimates in undetermined locations. Linyphiidae. — Meioneta fillmorana (Chamberlin 1919) is also primarily an arboreal spider. Its numbers peaked in May, after which there was a four month decline and then a marked rebound in October. Such a late season increase was unusual for spiders in this study especially since the October population of M. fillmorana consisted entirely of adults (Fig. 2). Six years of beat tray samples all showed a similar pattern, however. The species appears to be univoltine. Although we obtained virtually no overwintering data for M. fillmorana , it seems reasonable to infer, given the preponderance of females in the October collections, that overwintering probably takes place in the egg stage. At least two species of Erigone occurred in the orchards: Erigone dentosa O. Pickard-Canrbridge 1894 and Erigone aletris Crosby & Bishop 1928. We were unable to separate the two species with certainty. Adults of both sexes were present from at least May to October and were most abundant in October and June (Table 4). Although Erigone were well represented in the trees, sweep samples for 1996 and 1997 yielded a majority (65%) of the specimens that were collected. Erigone is also common on the ground based on pitfall trap collections (Miliczky et al. 2000), and thus has a broad distribution among habitats within the orchard. Tables 1 and 2 have data for less common Linyphiidae. Other families. — Data for less commonly collected spiders in other families is summarized in Tables 1 and 2. Included in this group: Anyphaena pacifica (Banks 1896) (Anyphaenidae) was occasionally found in the trees and immatures utilized overwintering shelters; Tetragnatha laboriosa Hentz 1850 (Tetragnathidae) was commonly swept from understory vegetation, and immatures were taken with some regularity in late season beat trays; Dictyna coloradensis Chamberlin 1919 (Dictynidae) was common only at the USDA research farm where it constructed webs on apple and pear leaves and tall weeds in adjacent uncultivated ground. DISCUSSION Many temperate zone spiders have a single generation per year (Gertsch 1979; Foelix 1996), and this appeared to be true of Washington species for which sufficient data were obtained. Within the broad latitudinal range of the temperate zone, however, factors that may influence spider development vary widely, and other life history patterns have been documented. About half of the 52 species studied in Denmark by Toft (1976, 1978) were biennial, and Almquist (1969) found a similar proportion of biennial species among 20 studied in Sweden. Almquist (1969) also observed that in Sweden the life cycles of some species were twice as long as the life cycles of the same or related species in southwestern Europe. Similarly, Philodromus cespitum is biennial in Nova Scotia (Dondale 1961) but univoltine farther to the south in Ontario (Putman 1967). In North America and Denmark maturation times among univoltine spiders form a continuum from the spring to the fall (Toft 1976; Gertsch 1979). Some litter inhabiting species of Linyphiidae mature and reproduce even during the winter months (Duffey 1956; Schaefer 1977). Washington species for which we acquired sufficient data appeared to have well- defined periods of reproduction (stenochrony) and could be classified as stenochronous with reproduction in spring and summer (Schaefer 1977). We noted that at a given time during the season, and also in the overwintering shelters, some species were represented by several developmental stages, whereas others were represented by only a few. Small, medium, and large immature Pelegrina aeneola could be found during much of the season (Fig. 1) and all these stages also overwintered (Table 1). Cheiracanthium mildei showed a similarly broad range of overwintering stages, while Phanias watonus was even more extreme as adults of both sexes also commonly overwintered (Table 1). In contrast, small immature Philodromus cespitum dominated the orchard collections from August to October (Fig. 1) and was the principal overwintering stage (Table 1). Philodromus calif ami- cus and Dictyna coloradensis spent the winter primarily as large immatures and/or penultimate males and penultimate females (Table 1). Factors tending to increase the length of time during the season when a given developmental stage is present include long-lived females that remain with an egg sac until the young disperse and produce more than one dutch of eggs. Egg sac guarding by P. aeneola was observed in the field, egg sacs were found as late as September, and females are capable of producing a second clutch of eggs. Pelegrina galatea (Walckenaer 1837) also guarded its eggs and produced multiple clutches in the laboratory (Horner & Starks 1972). We observed female C. mildei guarding eggs in the field and multiple clutches of eggs were produced under laboratory conditions (Mansour et al. 1980b). A number of the more common orchard spiders were poorly or not at all represented in overwintering shelters and presumably seek alternative sites inside or outside the orchard. Pekar (1999c) noted a similar phenomenon. Oxyopes scalaris, the only member of the family that occurs in Washington (Crawford 1988), presented an interesting case. Large numbers of small immature O. scalaris appeared in the orchards rather abruptly in July, and size increase in the population was observed as the season progressed. Large immatures and adults were rare in orchard collections (Fig. 1), however, and only two immatures were found in overwintering shelters (Table 1). The fate of the medium sized immatures that are so MILICZKY ET AL.— ORCHARD SPIDERS 571 Theridion neomexicanum Meioneta fillmorana Figure 2. — Percentage of adults and immatures of two common spider species found by month in combined beating tray and sweep net collections. F = female; M = male; sM = penultimate male; ssM = antepenultimate male (M. fillmorana only); 1mm - all other immature stages. Number at the top of the column is the total number of specimens taken for the month. common in the orchards in October remains to be determined. The opposite situation was noted in Steatoda hespera which utilized overwintering shelters but was not taken by beating tray or sweep net. Like most members of the genus, S. hespera webs are probably situated in rock and bark crevices and in cavities near the ground (Levi 1957) and would not have been sampled. Horton et al. (2001) changed a portion of their cardboard bands on pear and apple trees weekly from August to December. They determined that several species utilized the bands as temporary refuges during the autumn but overwin- tered elsewhere. Despite their abundance in terrestrial habitats and their exclusively predatory habits, Debach & Rosen (1991) noted a general neglect of spiders as potential biological control agents and attributed this, in part, to their generalist predatory habits. Other authors, noting the diversity of prey capture strategies and microhabitat exploitation patterns of spiders, have emphasized the contribution of the spider community as a whole to insect control in agroecosystems (Reichert & Lockley 1984; Marc & Canard 1997). Interest in the composition of spider faunas in orchards began over 50 years ago and appears to have increased given the number of studies conducted in recent years. However, studies attempting to evaluate the importance of spider predation on orchard pests are few (e.g., MacLellan 1973; Mansour et al. 1980a; Amalin et al. 2001; Miliczky & Calkins 2002). Table 4. — Summary by month of combined beating tray and sweep net collections of Erigone spp. Spider stages: ssM = antepenultimate male; sM = penultimate male. Month of Developmental stage collection ssM sM Male Female April 4 1 May 40 28 4 6 June 19 59 71 53 July 49 68 23 14 August 24 35 23 14 September 43 50 34 22 October 2 32 192 197 During this study we observed most of the common orchard spiders feeding on pests. Pelegrina aeneola , O. scalaris, P. cespitum , X. cunctator , and M. lepidus all used a variety of smaller pest species as prey, including leafhoppers, leafminers, aphids, thrips, and mites. The webs of M. fillmorana snared aphids and thrips as well as tiny flies and parasitoid wasps. The large salticid Phidippus clams took prey up to the size of an adult earwig. Some of the orchard spiders or a close relative may be important predators in agroecosystems more general- ly. Pelegrina aeneola and other members of the genus may be important in biological control because they are often abundant and are known to feed on pest insects (Horner 1972; Jennings & Houseweart 1978; Mason & Paul 1988). Oxyopes salticus Hentz 1845, a close relative of 0. scalaris, is a dominant predator in row crops in the United States and an important predator of pest insects (Young & Lockley 1985). Cheiracanthium mildei was described by Wise (1993) as a potentially important biological control agent in a number of agroecosystems. Mansour et al. (1980a) determined that C. mildei was the most effective spider predator of a lepidopteran pest of apples in Israel. Miliczky & Calkins (2002) rated it as having the greatest potential as a predator of pest leafrollers in Washington orchards out of 1 1 species tested. The role of spiders in orchard pest control is of considerable interest given the current trend toward reduced use of broad spectrum insecticides, the large numbers of spiders often observed when pesticide use is decreased or eliminated, and the great diversity among orchard-inhabiting spiders in size, behavior, and prey-capture strategies. All of these factors suggest that spiders should have substantial potential for contributing to orchard pest control. Future studies should further document the importance of this interesting but often overlooked group of beneficial organisms in controlling pest species in orchards and other agricultural systems. ACKNOWLEDGMENTS Funding for this research was provided by an Initiative for Future Agriculture and Food Systems (IFAFS) grant and by grants from the Washington Tree Fruit Research Commis- sion. Rod Crawford (The Burke Museum, University of Washington, Seattle), Andrew Moldenke (Oregon State University, Corvallis), and G.B. Edwards (Florida State 572 THE JOURNAL OF ARACHNOLOGY Collection of Arthropods, Gainesville) assisted with spider identifications. Rod Crawford, Lerry Lacey (USDA - Yakima Agricultural Research Laboratory), and two anonymous reviewers read earlier drafts of this paper and made helpful comments for its improvement. Merilee Bayer assisted with orchard sampling and with sorting and rearing of spiders. We are especially grateful to the many Yakima County apple and pear growers who allowed ready access to their orchards for sampling and observation. LITERATURE CITED Aiken, M. & F.A. Coyle. 2000. Habitat distribution, life history and behavior of Tetragnatha spider species in the Great Smoky Mountains National Park. Journal of Arachnology 28:97-106. Almquist, S. 1969. Seasonal growth of some dune-living spiders. Oikos 20:392-408. Amalin, D.M., J.E. Pena & R. McSorley. 2001. Predation by hunting spiders on citrus leafminer, Phyllocnistis citrella Stainton (Lepi- doptera: Gracillariidae). Journal of Entomological Science 36:199-207. Bogya, S. & V. Marko. 1999. Effect of pest management systems on ground-dwelling spider assemblages in an apple orchard in Hungary. Agriculture, Ecosystems & Environment 73:7-18. Bogya, S., V. Marko & Cs. Szinetar. 1999a. Comparison of pome fruit orchard inhabiting spider assemblages at different geograph- ical scales. Agricultural and Forest Entomology 1:261-269. Bogya, S., V. Marko & Cs. Szinetar. 2000. Effect of pest management systems on foliage- and grass-dwelling spider communities in an apple orchard in Hungary. International Journal of Pest Manage- ment 46:241-250. Bogya, S., Cs. Szinetar & V. Marko. 1999b. Species composition of spider (Araneae) assemblages in apple and pear orchards in the Carpathian Basin. Acta Phytopathologica et Entomologica Hun- garica 34:99- 121. Bostanian, N.J., C.D. Dondale, M.R. Binns & D. Pitre. 1984. Effects of pesticide use on spiders (Araneae) in Quebec apple orchards. Canadian Entomologist 116:663-675. Brown, M.W., J.J. Schmitt & B.J. Abraham. 2003. Seasonal and diurnal dynamics of spiders (Araneae) in West Virginia orchards and the effect of orchard management on spider communities. Environmental Entomology 32:830-839. Chant, D.A. 1956. Predacious spiders in orchards in south-eastern England. Journal of Horticultural Science 31:35—46. Crawford, R.L. 1988. An annotated checklist of the spiders of Washington. Burke Museum Contributions in Anthropology and Natural History No. 5. 48 pp. Debach, P. & D. Rosen. 1991. Biological Control by Natural Enemies. Second edition. Cambridge University Press, Cambridge, UK. 456 pp. Dondale, C.D. 1956. Annotated list of spiders (Araneae) from apple trees in Nova Scotia. Canadian Entomologist 88:697-700. Dondale, C.D. 1958. Note on population densities of spiders (Araneae) in Nova Scotia apple orchards. Canadian Entomologist 90:111-113. Dondale, C.D. 1961. Life histories of some common spiders from trees and shrubs in Nova Scotia. Canadian Journal of Zoology 39:777-787. Dondale, C.D. 1966. The spider fauna (Araneida) of deciduous orchards in the Australian Capital Territory. Australian Journal of Zoology 14:1157-1192. Dondale, C.D., B. Parent & D. Pitre. 1979. A 6-year study of spiders in a Quebec apple orchard. Canadian Entomologist 111:377- 380. Duffey, E. 1956. Aerial dispersal in a known spider population. Journal of Animal Ecology 25:85-111. Epstein, D.L., R.S. Zack, J.F. Brunner, L. Gut & J.J. Brown. 2000. Effects of broad-spectrum insecticides on epigeal arthropod biodiversity in Pacific Northwest apple orchards. Environmental Entomology 29:340-348. Foelix, R.F. 1996. Biology of Spiders. Second edition. Oxford University Press, New York. 330 pp. Gertsch, W.J. 1979. American Spiders. Second edition. Van Nostrand Reinhold Company, Toronto. 274 pp. Horner, N.V. 1972. Metaphidippus galathea as a possible biological control agent. Journal of the Kansas Entomological Society 45:324-327. Horner, N.V. & K.J. Starks. 1972. Bionomics of the jumping spider Metphidippus galathea. Annals of the Entomological Society of America 65:602-607. Horton, D.R., E.R. Miliczky. D.A. Broers, R.R. Lewis & C.O. Calkins. 2001. Numbers, diversity, and phenology of spiders (Araneae) overwintering in cardboard bands placed in pear and apple orchards of central Washington. Annals of the Entomolog- ical Society of America 94:405-414. Hukusima, S. 1961. Studies on the insect association in crop field. XXI. Notes on spiders in apple orchards. Japanese Journal of Applied Entomology and Zoology 5:270-272. Jennings, D.T. & M.W. Houseweart. 1978. Spider preys on spruce budworm egg mass. Entomological News 89:183-186. Legner, E.F. & E.R. Oatman. 1964. Spiders on apple in Wisconsin and their abundance in a natural and two artificial environments. Canadian Entomologist 96:1202-1207. Levi, H.W. 1957. The spider genera Crustulina and Steatoda in North America, Central America, and the West Indies (Araneae, Theridiidae). Bulletin of the Museum of Comparative Zoology 1 17:365-424. MacLellan, C.R. 1973. Natural enemies of the light brown apple moth, Epiphyas postvittana, in the Australian Capital Territory. Canadian Entomologist 105:681-700. Mansour, F., D. Rosen & A. Shulov. 1980a. A survey of spider populations (Araneae) in sprayed and unsprayed apple orchards in Israel and their ability to feed on larvae of Spodoptera littoralis (Boisd.). Acta Oecologica/Oecologica Applicata 1:189-197. Mansour, F., D. Rosen & A. Shulov. 1980b. Biology of the spider Chiracanthium mildei (Arachnida: Clubionidae). Entomophaga 25:237-248. Marc, P. & A. Canard. 1997. Maintaining spider biodiversity in agroecosystems as a tool in pest control. Agriculture, Ecosystems & Environment 62:229-235. Mason, R.R. & H.G. Paul. 1988. Predation on larvae of Douglas-fir tussock moth. Orgyia pseudotsugata (Lepidoptera: Lymantriidae), by Metaphidippus aeneolus (Araneae: Saliticidae). Pan-Pacific Entomologist 64:258-260. McCaffrey, J.P. & R.L. Horsburgh. 1980. The spider fauna of apple trees in central Virginia. Environmental Entomology 9:247-252. McCaffrey, J.P., M.P. Parrella & R.L. Horsburgh. 1984. Evaluation of the limb-beating sampling method for estimating spider (Araneae) populations on apple trees. Journal of Arachnology 11:363-368. Merrett, P. 1967. The phenology of spiders on heathland in Dorset I. Families Atypidae, Dysderidae, Gnaphosidae, Clubioni- dae, Thomisidae and Salticidae. Journal of Animal Ecology 36: 363-374. Merrett, P. 1968. The phenology of spiders on heathland in Dorset. Families Lycosidae, Pisauridae, Agelenidae, Mimetidae, Theridi- idae, Tetragnathidae, Argiopidae. Journal of Zoology, London 156:239-256. Miliczky, E.R. & C.O. Calkins. 2002. Spiders (Araneae) as potential predators of leafroller larvae and egg masses (Lepidoptera: Tortricidae) in central Washington apple and pear orchards. Pan-Pacific Entomologist 78:140-150. MILICZKY ET AL. ORCHARD SPIDERS 573 Miliczky, E.R., C.O. Calkins & D.R. Horton. 2000. Spider abundance and diversity in apple orchards under three insect pest management programmes in Washington State, U.S.A. Agricul- tural and Forest Entomology 2:203-215. Olszak, R.W., J. Luczak, E. Niemczyk & R.Z. Zajac. 1992. The spider community associated with apple trees under different pressure of pesticides. Ekologia Polska 40:265-286. Pekar, S. 1999a. Side-effect of integrated pest management and conventional spraying on the composition of epigeic spiders and harvestmen in an apple orchard (Araneae, Opiliones). Journal of Applied Entomology 123:115-120. Pekar, S. 1999b. Effect of IPM practices and conventional spraying on spider population dynamics in an apple orchard. Agriculture, Ecosystems & Environment 73:155-166. Pekar, S. 1999c. Some observations on overwintering spiders (Araneae) in two contrasting orchards in the Czech Republic. Agriculture, Ecosystems & Environment 73:205-210. Pekar, S. & F. Kocourek. 2004. Spiders (Araneae) in the biological and integrated pest management of apple in the Czech Republic, journal of Applied Entomology 128:561-566. Putman, W.L. 1967. Life histories and habits of two species of Philodromus (Araneida: Thonrisidae) in Ontario. Canadian Ento- mologist 99:622-631. Reichert, S.E. & T. Lockley. 1984. Spiders as biological control agents. Annual Review of Entomology 29:299-320. Schaefer, M. 1977. Winter ecology of spiders (Araneida). Zeitschrift fur angewandte Entomologie 83:1 13-134. Specht, H.B. & C.D. Dondale. I960. Spider populations in New Jersey apple orchards. Journal of Economic Entomology 53:810-814. Stiles, G.J. & F.A. Coyle. 2001. Habitat distribution and life history of species in the spider genera Theridion , Rugalliodes, and Want ha in the Great Smoky Mountain National Park (Araneae, Theridi- idae). Journal of Arachnology 29:396-412. Toft, S. 1976. Life-histories of spiders in a Danish beech wood. Natura Jutiandica 19:5-40. Toft, S. 1978. Phenology of some Danish beech-wood spiders. Natura Jutiandica 20:285-304. Wise, D.H. 1984. Phenology and life history of the filmy dome spider (Araneae: Linyphiidae) in two local Maryland populations. Psyche 91:267-288. Wise, D.H. 1993. Spiders in Ecological Webs. Cambridge University Press, Cambridge, UK. 328 pp. Wisniewska, J. & R.J. Prokopy. 1997. Pesticide effect on faunal composition, abundance, and body length of spiders (Araneae) in apple orchards. Environmental Entomology 26:763-776. Young, O.P. & T.C. Lockley. 1985. The striped lynx spider, Oxyopes salticus (Araneae: Oxyopidae), in agroecosystems. Entomophaga 30:329-346. Manuscript received 9 May 2007, revised 21 February 2008. 2008. The Journal of Arachnology 36:574—582 Distribution of Geraeocormobius sylvarum (Opiliones, Gonyleptidae): Range modeling based on bioclimatic variables Luis E. Acosta: CONICET - Catedra de Diversidad Animal I, Facultad de Ciencias Exactas, Fisicas y Naturales, Universidad Nacional de Cordoba, Av. Velez Sarsfield 299, X5000JJC Cordoba, Argentina. E-mail: lacosta@com.uncor.edu Abstract. The potential distribution of the harvestman Geraeocormobius sylvarum Holmberg 1887 (Opiliones, Gonyleptidae, Gonyleptinae) from Argentina, Brazil and Paraguay, is modeled using the presence-only, GIS-based method Bioclim. The model was run on 2.5 min resolution climate layers using 19 derived bioclimatic variables. The bioclimatic profile of the species is described, and presumable limiting factors in each part of the range are discussed. Modeled distribution of G. sylvarum shows a remarkable correspondence to the Alto Parana Atlantic forest ecoregion, with a marginal presence around the Araucaria forests and in gallery/flood forests towards the Southwest. Results support the 650 km yungas-Mesopotamia disjunction, as previously proposed, and reveal that localities in northwestern Argentina have extreme values concerning seasonality parameters with remarkably decreased rainfall in winter. Evidence suggesting that the disjunct pattern may have been derived by antropic introduction is briefly discussed. Keywords: Neotropics, bioclim, potential distribution, ecological niche modeling, environmental envelope Harvestmen (Arachnida, Opiliones) are generally regarded as a well suited taxon for biogeographic studies (Ringuelet 1959; Giribet & Kury 2007). Two key features make them useful for those purposes: their low vagility and a close dependence on environmental conditions, mainly humidity (Acosta 2002; Pinto-da-Rocha et al. 2005; Machado et al. 2007). Distribution of most species is thus dependent on the geographic continuity of suitable environments (Acosta 2002). In addition, harvestman endemicity may be particularly remarkable in some areas (Machado et al. 2007). Small-area endemicity is striking in forested ranges like the Brazilian Serra do Mar (Pinto-da-Rocha et al. 2005), and is also important in the montane forests of northwestern Argentina, the so called “yungas” region (Acosta 2002). However, small-ranged endemics should not be seen as the rule for harvestman distribution. Under certain conditions, many species are known to spread over thousands of knr (Curtis & Machado 2007) as long as the suitable environment is not restricted by any geographic or ecological barrier. For example, while many central European species appear geo- graphically restricted, no less than 30 harvestmen in that region extend over quite large ranges (Martens 1978). Broad ranged species are also typical for most of the Argentinean Mesopo- tamia, which is the humid and sub-humid region between the Parana and Uruguay rivers (Acosta 2002). Its northernmost portion, roughly matching the administrative province of Misiones. bears subtropical forest physiognomy - actually a part of the “Paranense Biogeographic Province” (Cabrera & Willink 1973) that covers adjacent areas in Paraguay and Brazil as well. The rest of Mesopotamia is a mosaic of shrubs, swamps, grasslands and gallery forests (Hueck & Seibert 1972; Cabrera & Willink 1973). In accordance with these differences, the Mesopotamian opiliofauna has been split into two different, though overlapping, sub-areas (Acosta 2002): the Misiones sub- area and the Mesopotamian sensu strieto sub-area. Towards the West, as precipitation decreases, these humid and sub-humid realms give way to the semiarid Chaco, an effective distribu- tional limit for Mesopotamian harvestmen (Acosta 2002). Geraeocormobius sylvarum Holmberg 1887 (Opiliones, Gonyleptidae, Gonyleptinae) is a large and conspicuous harvestman known to inhabit subtropical forests in north- eastern Argentina, southern Brazil and southeastern Paraguay (Ringuelet 1959; Kury 2003; Acosta et al. 2007), being thus characteristic of the Misiones sub-area (Acosta 2002). Acosta et al. (2007) provided several new records for this species; among them three localities in the province of Tucuman, Argentina, revealing a disjunct presence in montane forests in northwestern Argentina (NWA). These separate populations lie about 650 km away from the westernmost record in the core area, with the sub-xeric Chaco in between. As stressed by Acosta et al. (2007), most records of G. sylvarum concentrate in the Alto Parana Atlantic forests ecoregion (referred to as “Paranense forests” below) and in some adjacent sectors of the Araucaria moist forest ecoregions (nomenclature after Olson et al. 2001). Argentinean captures outside the men- tioned ecoregions (in provinces of Corrientes and Chaco) seemingly are associated with gallery forests and/or seasonal inundation sites (the Humid Chaco and Southern Cone Mesopotamian savanna ecoregions of Olson et al. (2001)). In turn, findings of G. sylvarum in NWA correspond to the “tucumano-boliviano” forests, or “yungas” (Hueck & Seibert 1972; Acosta 2002), or the Southern Andean Yungas ecoregion (Olson et al. 2001). The available localities show some spatial bias, since they are concentrated in and around the province of Misiones (a traditionally well sampled area; Ringuelet 1959), while extensive areas in Paraguay are left almost undocumented (cf. Fig. 3). Interestingly, two further Mesopotamian harvestmen, Dis- cocyrtus dilatatus Sorensen 1884 and D. prospicuus (Holmberg 1876) (Gonyleptidae, Pachylinae) have Mesopotamia-yungas disjunct ranges as well (Acosta 1995, 2002). A very basic question remains unanswered, however: Are Chacoan condi- tions really inhospitable to Mesopotamian species or is our record incomplete, leading us to wrongly assume this region to be hostile for harvestmen? Nothing is known about the climatic tolerances of D. dilatatus, D. prospicuus or G. 574 ACOSTA— DISTRIBUTION MODELING OF GERAEOCORMOBIUS (OPILIONES) 575 sylvarum as very little is known about harvestmen physiology and tolerance to physical factors in general (Santos 2007). It is generally accepted that harvestman distribution is governed by climatic constraints, mainly humidity and temperature, but papers addressing how actual climatic conditions affect a given species are almost lacking (Curtis & Machado 2007; Machado et al. 2007). The occurrence records contain some useful clues. Point records are the very basis for discovering a species range, though the vast majority of Neotropical harvestmen have been scarcely recorded, so that species with more than 15 records are rare (cf. Kury 2003). In fact, most distributional patterns are then intuitively extrapolated from the few records available, emphasizing the need for gathering more data and filling in the gaps. Alternatively, recent developments aimed to model potential species ranges, using available records and several types of environmental predictors, offer innovative methods to help detect areas where presence of a given species may be expected though still not documented (Guisan & Zimmermann 2000; Guisan & Thuiller 2005; Hernandez et al. 2006; Peterson 2006). Using a Geographic Information System (GIS) and appropriate climate information, the envelope approach infers the species’ bioclimatic profile, a species- specific feature (actually, a subset of the fundamental ecological niche) that gives us a rough insight into their presumed tolerance limits, at least with respect to the variables used to build the model. It is then possible to project this profile onto the geographic space, to identify all areas with similar climatic conditions, i.e., where the species would meet a potentially suitable environment (Guisan & Zimmermann 2000; Hijmans & Graham 2006; Hernandez et al. 2006; Pearce & Boyce 2006). This paper is intended to characterize the bioclimatic profile of G. sylvarum and to model its potential distribution using the bioclim algorithm, taking advantage of the number of records now available for this species (close to the minimum sample size for the method to attain an acceptable accuracy; Hernandez et al. 2006). Thereby it is also intended to test if the disjunct Mesopotamian-yungas pattern is supported by the model and to identify areas in which future sampling efforts would be productive in filling in distribution gaps. Results are matched to major biomes and habitat types in order to get a more comprehensive understanding, though still preliminary, of the ecological requirements and potential range of this wide-ranged harvest- man. METHODS Data acquisition. -All published references for G. sylvarum were taken into account (Holmberg 1887; Ringuelet 1959; Soares & Soares 1985; Kury 2003; Acosta et al. 2007). Historical records were inspected for taxonomic reliability (cf. Acosta et al. 2007 for localities excluded), while three localities from the province of Misiones (published by Ringuelet 1959) remained unrecognizable and were set aside: “Campamento Yacu-Poi, near Puerto Bemberg”; “60 km Puerto Iguazu” (which direction?), and “Pasarela Rio Uruguay.” The full dataset consisted of 48 unique point localities; during the analysis, duplicate records from the same gridcell were removed by the software, resulting in 46 effective records. Localities were identified and geo-referenced using printed road maps and digital gazetteers (mainly NGA GEOnet Names Server [GNS], United States Board on Geographic Names; Google Earth ®), all cross-checked for final accuracy. The label data or collector’s information was used to determine location as precisely as possible. Published records typically refer to a locality, which may represent not more than the nearest reference to the actual collecting site; such an imprecision in coordinates (impossible to measure) is deemed not to affect the results considering both the regional scale and coarse approach used. Climate layers. — The model was run on Worldclim 1.4. (Hijmans et al. 2005a), a set of global climate layers, containing extrapolated monthly data for the 1950-2000 period on maximum, minimum and mean temperature, and precipitation. The 2.5 min resolution (i.e., approximately 4.5 x 4.5 km gridcell) was selected. Information contained in Worldclim (climate data, together with a digital elevation model) is used by the software to derive the 19 bioclimatic variables available for modeling, listed in Table 1. The abbreviation “be” followed by a number is used below to identify each of these bioclimatic variables. Modeling method. -Modeling was performed through the presence-only method Bioclim, as implemented in Diva-Gis 5.4 (Hijmans et al. 2005b). Bioclim is a frequency distribution- based algorithm, which calculates the envelope that bounds the bioclimatic preferences of the species (Fischer et al. 2001; Walther et al. 2004; Hernandez et al. 2006). Values of each derived bioclimatic variable are extracted from all localities and arranged in a cumulative frequency distribution (cf. Fig. 2). The envelope is defined as a multi-dimensional hyper- box (each variable representing a dimension) delimiting, at a given percentile, the climatic conditions in the occurrence localities (Guisan & Zimmermann 2000). This set of values constitutes the bioclimatic profile of the species (Fischer et al. 2001), as summarized for the target species in Table 2. In the potential distribution maps (Figs. 1, 4-5), gridcells are scored as suitable (if within the envelope; i.e., the presence of the species can be expected) or unsuitable (outside the envelope). User-defined percentiles were set to define the extent of the envelope (as cut-off) or to rank the gridcells suitability (Fischer et al. 2001; Hijmans et al. 2005b; Hernandez et al. 2006). Evaluation. -The accuracy of the predictive range generated by the model was assessed by calculating the AUC (area under curve) in a receiver operating characteristic (ROC) plot, and the maximum Kappa (rnax-K) value, both analyses made in Diva-Gis (Hijmans et al. 2005b). In this study, 70% of the original points were randomly resampled as training data to perform 20 repetitions of the model. Test data included pseudo-absence points selected at random from the back- ground. ROC/max-K were calculated against the grids stack (0-100 percentile) of the 20 range polygons that were modeled using the training points. AUC values over 0.8 are deemed to reflect a “good” model performance; above 0.9 the accuracy is considered “high” (Luoto et al. 2005). In turn, max-K over 0.4 are deemed to be “good” and “excellent” if above 0.75 (Randin et al. 2006). Input variants. — A separate run was performed with records from NWA removed to verify if the species is still predicted in that area. To identify the factors limiting the distribution, the 576 THE JOURNAL OF ARACHNOLOGY Table 1. — Median, minimum and maximum values, and range for all 19 biocliinatic variables in the envelope of Geraeocormobius sylvarum. Absolute temperature values are in degrees Celsius (° C), precipitation in mm. Numbers preceding each variable are referred to in the text and Table 2. Main differences between profiles with and without yungas records are emphasized in bold. Biocliinatic variables All records (n = 46) Yungas removed (n = 44) Median Min-max (range) Median Min-max (range) ( 1 ) Annual mean temperature 20.39 16.46-23.12 (6.66) 20.62 16.46-23.12 (6.66) (2) Mean monthly T° range 12.43 10.65-13.83 (3.18) 12.43 10.65-13.83 (3.18) (3) Isothermality (2/7 x 100) 56.08 47.25-65.00 (17.75) 56.21 49.48-65.00 (15.52) (4) T seasonality (STD x 100) 368.87 252.92-461.59 (208.67) 366.90 252.92-430.76 (177.84) (5) Max T° of warmest month 31.65 27.20-34.20 (7.00) 31.70 27.20-34.20 (7.00) (6) Min T° of coldest month 9.25 4.80-12.50 (7.70) 9.40 4.80-12.50 (7.70) (7) T° annual range (5-6) 22.50 19.00-25.40 (6.40) 22.45 19.00-24.30 (5.30) (8) Mean T° wettest quarter 21.52 16.20-25.23 (9.03) 21.36 16.20-25.23 (9.03) (9) Mean T° driest quarter 16.84 13.40-19.22 (5.82) 16.92 14.03-19.22 (5.18) (10) Mean T° warmest quarter 24.76 20.33-27.58 (7.25) 24.89 20.33-27.58 (7.25) (11) Mean T° coldest quarter 16.01 12.35-18.50 (6.15) 16.03 12.50-18.50 (6.00) (12) Annual precipitation 1723.5 934-2235 (1301) 1730.5 1250-2235 (985) (13) Precipitation wettest month 188.5 157-245 (88) 186 157-245 (88) (14) Precipitation driest month 99.5 11-150 (139) 101 42-150 (108) (15) Precipitation seasonality (CV) 20.34 9.42-83.74 (74.32) 19.94 9.42-41.88 (32.46) (16) Precipitation wettest quarter 495.5 432-625 (193) 495.5 432-625 (193) (17) Precipitation driest quarter 344.5 40-496 (456) 349.5 140-496 (356) (18) Precipitation warmest quarter 457 354-625 (271) 455 354-625 (271) (19) Precipitation coldest quarter 360 40-540 (500) 362 140-540 (400) model (full dataset) was alternatively run for each variable separately and combining some of them so as to visually inspect the effects on the resulting predicted range. RESULTS Biocliinatic profile.-- Table 1 summarizes values relevant to the biocliinatic tolerance range of G. sylvarum. Separate profiles with and without the Tucuman localities are given from which it is clear that, for some variables, conditions in the yungas suggest differences from the core area (cf. Table 2 also). Sites from Tucuman represent both the westernmost records and the highest elevation. In those localities precip- itation (be 12) is the lowest (Fig. 2B). with rainfall decreasing substantially during the winter (Fig. 2D). There, seasonality is higher than in sites of the core area (Figs. 2A, C); in particular, precipitation seasonality (be 15) is strongly skewed to low values, with both Tucuman records clearly separated by a decided gap from the rest (Fig. 2C). When these localities of NWA are set aside, Rio Tragadero and 10 km Puerto Antequera (close to each other) become the species’ western- most edge and come to hold many of the biocliinatic extremes related to seasonality and winter decrease of precipitation (Table 2 and Fig. 3). However, geographically extreme localities did not necessarily hold the highest or lowest biocliinatic values in all cases (Table 2). Other localities with many biocliinatic variables showing extreme values are Clevelandia (Santa Catarina, Brazil) and Asuncion (Para- guay). Clevelandia has nine end values indicating, in general, a cooler and more humid climate than the rest; in turn, Asuncion stands as the warmest site of the species range, with highest values for six variables, all temperature-related. With respect to the resulting envelope, a permissive percentile cut-off of 0.005 allowed 33 out of 46 observations (71.7%) to be included within all possible 171 bidimensional envelopes (the remaining 13 observations being outsiders in at least one bidimensional envelope). With the default percentile of 0.025, localities within the species envelope are reduced to 21 out of 46 observations (45.7%). Potential range. — The predicted range of G. sylvarum under the model parameters is displayed in Fig. 1. To a great extent, the predicted core area of G. sylvarum roughly matches the Paranense forests ecoregion (Fig. 3). In eastern Paraguay, where records are almost lacking, the prediction partially redraws the boundaries of this ecoregion with the contiguous Humid Chaco (Fig. 3). On the Brazilian side, the predicted range seemingly enters the Araucaria forests only marginally, i.e., in areas surrounded by complex eastward Paranense projections that follow large rivers. A large portion of the Araucaria ecoregion is scored as unsuitable; moreover, it is to be noted that Clevelandia, one of the localities with many biocliinatic extreme values (Table 2), is placed near this presumable distributional limit (Fig. 3). In Brazil the modeled range reaches up to the southern states of Mato Grosso do Sul and Sao Paulo in the North (though with no records so far), and weakly up to the western slopes of the Serra do Mar in the East (Fig. 1). Geraeocormobius sylvarum was not hitherto collected in the eastern slopes of the Serra do Mar; probably replaced there by its congener Geraeocormobius rohri (Mello- Leitao 1933) (A.B. Kury, pers. comm.). On the other side, occurrences of G. sylvarum in the northern province of Corrientes and eastern Chaco give support to the potential areas southwest of the paranense forests. Subtropical vegeta- tion partially extends into northern Corrientes, as well as along flood and gallery forests along the Parana and Paraguay rivers (Hueck & Seibert 1972), and this seems to provide suitable conditions for G. sylvarum some hundreds of kilometers away from the main range. With the full dataset, the biocliinatic model predicts the presence of G. sylvarum in the yungas but not across the sub- xeric Chaco (Fig. 1); thus, the presumed disjunction is ACOSTA— DISTRIBUTION MODELING OF GERAEOCORMOBIUS (OPILIONES) 577 C/5 ”0 53 5 G dJ 3 •§ c g w '5b C/5 C o •r o CJ g X > w dJ O X *55 4J ,p e x "p S o c .5 G O w Oh C/5 (D rt ^ g .a .G c/5 G .G o — ■G G 'S ° a o .O -o 2 "O 03 £ ^ 03 a g Q Q> b/j P T3 3 « dJ o "§ s Z, dJ •— S-H G P o 3 <3 w c G kG c^J g 0) G *0 o .G H ° j s S5J P *03 ^ c+h iG C/5 G S ^ g g g on r: ( G G G G G — 1 G o (U 1) l) (U , jj G 03 c/5 G G ^ O > > O CJ 03 03 CJ < H H < I ^ I I I _ os O °°' — r~ ‘too'5 I (f| I | „ | Os Os PI ^ X 1 4 so , — 1 in r— ' : pT os os os p ; Xh X X pf ^ 2 ^ rn w Os" w p O Pj Pi P- P- 00 O 00 cp cn s© o o o m 00 00 00 00 cn iri SO so SO P~ P- QC PI P) PI PI PI PI I I I I I I - „ ^ o ^ ■ !b S2 CJ S 00 1 1 1 rt,2 1 1 bi . r 1 so ^ in pi n p- o p- p- p- 0 0 so O X X X o o X O X X ^ o o m 00 o cp os cn 'rt p so p sd in n so pi pi pi pi pi pi pi I I I I I I I O m m o O O p o cn m o o o x o m »n »n o in 00 m m — o pi *n in cn in Tf p p^ p p pi pi pi pi pi pi pi I I I I I I I m p cn p o m o cn cn p p cn p p p p p o o p p QC pi X o cn o cn cn X X cn X X X X o o p X i n X cn X o X o cn cn X p X X X X pi o in p X en cn cn X X x SO o X X X so cn X p p 0© X pi in p it] in in 0© oo p «n in X X X p pi pi o as pi cn pi pi X X X ITi in in »n X in •n in •n «n in in >n >n »n «n g on 2 o I «S c2 tx U G G G C^0 C/0 00 G -0 G G vG X G G H H G VG S o G CJ C/5 C/5 C/5 C/5 C/5 CJ b CJ 1) aj G g O H C/5 C/5 C/5 C/5 G t=j 5 s u 'I % i i i £ £ G G - - u „ „ - g c g a a a G G *-* 1 G O O O G G G X X X 0- cu 00 a! oC oi GGGGGGGGGGG .= .5 .5 .5 .= .2 .2 .5 .£ .2 .2 3 J— « S— « <<<<<<<<<< i ♦♦♦♦* *♦♦♦ J r L £0 04 13 E ZJ 1 ♦* J / ♦ £ i i i i i i i O 0.2 0.0 ‘ 1 L i_i 1 * 1 , l , O 10 20 30 40 50 60 70 80 90 bc15: precipitation seasonality D O 100 200 300 400 bc17: precipitation driest quarter Figure 2. — Bioclimatic profile of Geraeocormobius sylvarum, selected variables: full data set plotted for cumulative relative frequency; arrows indicate the position of the two localities from province of Tucuman (NWA). A: bc4 - temperature seasonality (standard deviation x 100); B: be 12 annual precipitation (mm); C: bcl5 - precipitation seasonality (coefficient of variation); D: bcl7 - precipitation of the driest quarter (mm). ACOSTA- - DISTRIBUTION MODELING OF GERAEOCORMOBIUS (OPILIONES) 579 Figure 3. — Correspondence of the modeled range of Geraeocormobius sylvarum (grey area, 0-100 percentiles) and the Alto Parana Atlantic forests ecoregion (squared pattern fill); the overlap is shown as a lighter grey squared area. AMF: Araucaria moist forests (bordered by a thin line), where G. sylvarum is predicted to occur marginally. Dots: all locality records for the species. References of selected localities: 1 : Asuncion; 2: Rio Tragadero; 3: Clevelandia; 4: Caviuna; 5: Paranavai. effects in the core area: just a small clipping occurs, reducing the range on marginal sides, like the Corrientes-Chaco portions, and its northernmost extension in Brazil. Limiting factors. — Models of G. sylvarum run separately, with either temperature ( be 1 — 11) or precipitation (be 1 2—19) variables, uncover the relative contribution of these variables to the predicted range shape. Prediction based on temperature variables bcl-bcll (Figs. 4B, 5A) more closely recovers the model obtained by all 19 variables together, with a slight “permissiveness” in the southernmost border (province of Corrientes) and especially in NWA (patchy suitable areas appear along the yungas in Tucuman, Jujuy, Salta, and southern Bolivia). Precipitation variables bcl2-bcl9 (Figs. 4C, 5B) clip most of these areas away from the model, but, in turn, extend suitable conditions considerably eastwards and north- wards. In fact, many precipitation variables (when run alone) predict G. sylvarum in the yungas, but almost all prevent it entering the Chaco, some being more rigorous (be 15, be 16, be 18), others more “permissive” (bcl2, be 14, bcl7, bcl9); only be 13 enables a somewhat continuous range. In contrast, only a few temperature variables (bc5 and bc8) contribute to the Chaco gap. Temperature variables are more related to range restrictions in the North and the East. The northernmost boundaries in Brazil, where the range meets the Cerrado and is limited by high temperatures, are fairly well shaped by bcl, bc4, bc6, bc9, and bell. As noted above, the modeled range Figure 4. — Detail of the predictive modeling in province of Tucuman, Argentina. A: model with all 19 bioclimatic variables (small dots indicate the occurrence records), B: model with only temperature variables (bcl— be 1 1), C: model with only precipitation variables (be 1 2 be 19). Light grey: 0-25 percentile; dark grey: above 0.25. 580 THE JOURNAL OF ARACHNOLOGY Figure 5. — Potential distribution of Geraeocormobius sylvarum modeled only with temperature or precipitation variables (light grey: 0-25 percentile; dark grey: above 0.25). A: modeled range with temperature variables (be 1 —be 1 1), B: modeled range with precipitation variables (bcl 2— be 19). seems to circumvent the Araucaria forest, entering this ecoregion only marginally (Fig. 3). Several temperature variables (bcl, bc5, bc6, bc8, bc9, bclO, bell) leave a defined gap there. The unsuitability appears to be related to the high elevation, which in turn causes the seemingly unsuitable cold climate: Clevelandia, elev. 890 m, is the highest record of G. sylvarum in Brazil and bears several extreme cold climate values (Table 2), but many negative sites in the Araucaria area are above 1 100 m. Accuracy. — Modeling with the training samples (30% of the original points randomly removed) consistently recovered the main portions of the range predicted with all points. More than half of the 20 repetitions matched in most parts of the core area, and just marginal sectors overlapped in only 10 repetitions or less. AUC values ranged from 0.8505 to 0.9402 (mean = 0.8935), and max-K from 0.7082 to 0.8832 (mean = 0.7900); thus, individual and average AUC and max-K values resulted in good to high model performances. These values are consistent ACOSTA— DISTRIBUTION MODELING OF GERAEOCORMOBIUS (OPILIONES) 581 with performances of Bioclim obtained by Hernandez et al. (2006) for similar sample sizes, and with comparative evalua- tions made by Sangermano & Eastman (2007). DISCUSSION As stressed by Peterson (2006), for the vast majority of species nothing more is known than a few “dots on maps” - and certainly, this applies for most Neotropical harvestmen, too. Ecological niche (or habitat suitability) modeling offers a first step towards inferring the basic ecological dimensions that are relevant to limit the species’ distribution. While previous knowledge just imprecisely related G. sylvarum to Paranense forests (and thereby assumed it was dependent on humid conditions), the bioclimatic analysis provided, for the first time, defined values to describe and discuss the species profile. The envelope method proved to be well suited as an initial approach when species records and biological knowl- edge are still scarce (Pearce & Boyce 2006). Results were convincing when matching the predicted range polygon to the Paranense forests ecoregion {Fig. 3), a biogeographical area that was previously associated with G. sylvarum based on a “non analytical, expert-based” assessment (Acosta et al. 2007). The model accuracy as measured by AUC and max-K was acceptable as well. One of the main issues tackled in this paper, the Mesopotamia-yimgas disjunction, received support from this model; but at the same time, some new questions arose concerning the presence of G. sylvarum in NWA. Despite the availability of presence records, the species is only weakly predicted there. This contrasts with other Mesopotamian harvestmen (D. dilatatus and D. prospicuus among them) that, in preliminary models, were predicted in the yungas, with or without positive records in NWA (Acosta 2007). Both D. dilatatus and D. prospicuus belong to the Mesopotamian sensu stricto sub-area (Acosta 2002), so their climatic requirements are not expected to be as humidity-dependent as for G. sylvarum. Since Bioclim is deemed to over-predict a species’ range (Peterson 2001), these facts, together with the profile values for some variables, suggest an important disparity of the bioclimatic conditions of NW and NE Argentina concerning preferences of G. sylvarum. However, one should not lose sight of the importance of the resolution of the climatic data. In general the resolution used (2.5 min) is acceptable at a regional scale, but its coarseness may not adequately reflect small-distance variations in montane areas, like in Tucuman (records range from 500 to 1250 m elevation within a distance of less than 3 km; i.e., less than the gridcell size). As for the causes of the disjunction, a satisfactory explanation remains an open question. Acosta (1995, 2002) suggested that the disjunct ranges of the two Discocyrtus species may be a consequence of paleoclimatic cycles, with associated expansion/retraction events affecting the humid forests. A hypothetical vegetational “bridge” (as proposed by Nores 1992 for birds) would have acted as a corridor, enabling Mesopotamian harvestmen to expand their ranges up to the yungas, to leave isolated populations there as climate turned rigorous and the forests retracted. In the case of G. sylvarum , however, several elements may render this explanation less likely. Records in the yungas are actually few, indicating a quite limited distribution, and the bioclimatic model accord- ingly predicts this species only weakly for the region. The yungas is one of the best sampled areas in Argentina for harvestmen (Acosta 2002), so it seems unlikely to find G. sylvarum elsewhere in NWA. People from the species’ core range are in general aware of G. sylvarum, probably because of its large size, abundance and strong odor. These conspicuous and smelly harvestmen are also locally well known in El Corte, although people that have lived there for the last 30 years suggest that the species “appeared” only in recent years (A. M. Frias, pers. comm.). These observations, together with the apparent tolerance of G. sylvarum to disturbed areas (Acosta et al. 2007), may suggest that transportation by humans in historic times might best account for the presence of this Mesopotamian harvestman in Tucuman. ACKNOWLEDGMENTS I am indebted to Maria Laura Juarez, Gustavo Flores, Dardo Marti, Adriano Kury, and Miryam Damborsky for providing coordinates or details of the collecting sites of G. sylvarum. Comments and suggestions given by Raquel M. Gleiser, Soren Toft, Paula Cushing and an anonymous referee greatly improved the manuscript. Financial support was given by Consejo Nacional de Investigaciones Cientificas y Tecnicas (CONICET: P.I.P. N° 6319) and SECyT-UNC. The author is a researcher of CONICET. LITERATURE CITED Acosta, L.E. 1995. Nuevos hallazgos de Discocyrtus dilatatus en Argentina, con notas sobre taxonomia, sinonimia y distribucion (Opiliones, Gonyleptidae, Pachylinae). Revue Arachnologique 10: 207-217. Acosta, L.E. 2002. Patrones zoogeograficos de los Opiliones argentinos (Arachnida: Opiliones). Revista Iberica de Aracnologia 6:69-84. Acosta, L.E. 2007. Distribution of harvestmen (Opiliones) in the Argentinean Mesopotamia: a modeling approach based on bioclimatic variables. 17th International Congress of Arachnology, Sao Pedro, SP, Brazil. Abstracts, p. 49. Acosta, L.E., A.B. Kury & M.L. Juarez. 2007. New records of Geraeocormobius sylvarum (Arachnida, Opiliones, Gonyleptidae), with a remarkable disjunction in northwestern Argentina. Boletin de la Sociedad Entomologica Aragonesa 41:303-306. Cabrera, A.L. & A. Willink. 1973. Biogeografia de America Latina. Coleccion Monografias Cientificas O.E.A., serie Biologia, n° 13, pp. I-VI, 1-117. Curtis, D.J. & G. Machado. 2007. Ecology. Pp. 280-308. In Harvestmen: The Biology of Opiliones. (R. Pinto-da-Rocha, G. Machado & G. Giribet, eds.). Harvard University Press, Cam- bridge, Massachusetts. Fischer, J., D.B. Lindenmayer, H.A. Nix, J.L. Stein & J.A. Stein. 2001. Climate and animal distribution: a climatic analysis of the Australian marsupial Trichosurus caninus. Journal of Biogeogra- phy 28:293-304. Giribet, G. & A.B. Kury. 2007. Phylogeny and biogeography. Pp. 62-87. In Harvestmen: The Biology of Opiliones. (R. Pinto- da-Rocha, G. Machado & G. Giribet, eds.). Harvard University Press, Cambridge, Massachusetts. Guisan, A. & W. Thuiller. 2005. Predicting species distribution: offering more than simple habitat models. Ecology Letters 8:993- 1009. Guisan, A. & N.E. Zimmermann. 2000. Predictive habitat distribu- tion models in ecology. Ecological Modelling 135:147-186. 582 THE JOURNAL OF ARACHNOLOGY Hernandez, P.A., C.H. Graham, L.L. Master & D.L. Albert. 2006. Tire effect of sample size and species characteristics on perfor- mance of different species distribution modeling methods. Eco- graphy 29:773-785. Hijmans, R.J., S.E. Cameron, J.L. Parra, P.G. Jones & A. Jarvis. 2005a. Very high resolution interpolated climate surfaces for global land areas. International Journal of Climatology 25:1965-1978. Hijmans, R.J. & C.H. Graham. 2006. The ability of climate envelope models to predict the effect of climate change on species distributions. Global Change Biology 12:1-10. Hijmans, R.J., L. Guarino, A. Jarvis, R. O’Brien, P. Mathur, C. Bussink, M. Cruz, I. Barrantes & E. Rojas. 2005b. DIVA-GIS, version 5.4.0. 1. Online at http://www.diva-gis.org/ Holmberg, E.L. 1887. Viaje a Misiones. Boletin de la Academia Nacional de Ciencias, Cordoba 10:5-391. Hueck, K. & P. Seibert. 1972. Vegetationskarte von Siidamerika. Vegetationsnronographien der einzelnen GroBraumen, 2a: 1-71 + map. G. Fischer, Stuttgart. Kury, A.B. 2003. Annotated catalogue of the Laniatores of the New World (Arachnida, Opiliones). Revista Iberica de Aracnologia, Volumen especial monografico 1:5-337. Luoto, M., J. Pdyry, R.K. Heikkinen & K. Saarinen. 2005. Uncertainty of bioclimate envelope models based on the geo- graphical distribution of species. Global Ecology and Biogeogra- phy 14:575-584. Machado, G., R. Pinto-da-Rocha & G. Giribet. 2007. What are harvestmen? Pp. 1-13. In Harvestmen: The Biology of Opiliones. (R. Pinto-da-Rocha, G. Machado & G. Giribet, eds.). Harvard University Press, Cambridge, Massachusetts. Martens, J. 1978. Weberknechte, Opiliones. Die Tierwelt Deutsch- lands 64:1—464. Gustav Fischer, Jena Nores, M. 1992. Bird speciation in subtropical South America in relation to forest expansion and retraction. Auk 109:346-357. Olson, D.M., E. Dinerstein, E.D. Wikramanayake, N.D. Burgess, G.V.N. Powell, E.C. Underwood, J.A. D’Amico, H.E. Strand, J.C. Morrison, C.J. Loucks, T.F. Allnutt, J.F. Lamoreux, T.H. Ricketts, I. Itoua, W.W. Wettengel, Y. Kura, P. Hedao & K. Kassem. 2001. Terrestrial ecoregions of the world: a new map of life on Earth. BioScience 51:933-938. Pearce, J.L. & M.S. Boyce. 2006. Modelling distribution and abundance with presence-only data. Journal of Applied Ecology 43:405-^412. Peterson, A.T. 2001. Predicting species’ geographic distributions based on ecological niche modeling. Condor 103:599-605. Peterson, A.T. 2006. Uses and requirements of ecological niche models and related distributional models. Biodiversity Informatics 3:59-72. Pinto-da-Rocha, R., M.B. da Silva & C. Bragagnolo. 2005. Faunistic similarity and historic biogeography of the harvestmen of southern and southeastern Atlantic Rain Forest of Brazil. Journal of Arachnology 33:290-299. Randin, C.F., T. Dirnbock, S. Dullinger, N.E. Zimmermann, M. Zappa & A. Guisan. 2006. Are niche-based species distribution models transferable in space? Journal of Biogeography 33:1689- 1703. Ringuelet, R.A. 1959. Los aracnidos argentinos del Orden Opiliones. Revista del Museo Argentine de Ciencias Naturales, Ciencias Zoologicas 5:127-439. Sangermano, F. & R. Eastman. 2007. Linking GIS and Ecology - The use of Mahalanobis typicalities to model species distribution. Pp. 1-13. In Memorias XI Conferencia Iberoamericana de Sistemas de Informacion Geografica. (G.D. Buzai, ed.). Buenos Aires. Santos, F.H. 2007. Ecophysiology. Pp. 473-488. In Harvestmen: The Biology of Opiliones. (R. Pinto-da-Rocha, G. Machado & G. Giribet, eds.). Harvard University Press, Cambridge, Massachu- setts. Soares, B.A.M. & H.E.M. Soares. 1945. Alguns opilides do Museu Nacional do Rio de Janeiro. Papeis do Departainento de Zoologia do Estado de Sao Paulo 5:221-226. Soares, H.E.M. & B.A.M. Soares. 1985. Contribution a 1’etude des opilions (Opiliones: Cosmetidae, Phalangodidae, Gonyleptidae) du Paraguay. Revue Suisse de Zoologie 92:3-18. Tavares, M.L.R. 1980. Novas ocorrencias de opiiioes no Rio Grande do Sul e descrigao da femea de Melloleitaniana riodariensis Soares & Soares, 1945 (Arachnida-Opiliones-Gonyleptidae). Iheringia, Serie Zoologia 55:155-159. Walther, B.A., M.S. Wisz & C. Rahbek. 2004. Known and predicted African winter distributions and habitat use of the endangered Basra reed warbler ( Acrocephalus griseldis) and the near-threatened cinereous bunting ( Emberiza cineracea). Journal of Ornithology 145:287-299. Manuscript received 1 June 2007, revised 18 February 2008. 2008. The Journal of Arachnology 36:583-594 Ecology and web allometry of Clitaetra irenae , an arboricolous African orb-weaving spider (Araneae, Araneoidea, Nephilidae) Matjaz Kuntner1’2’5, Charles R. Haddad1, Gregor Aljancic4 and Andrej Blejec4: 'Institute of Biology, Scientific Research Centre of the Slovenian Academy of Sciences and Arts, Novi trg 2, SI- 1001 Ljubljana, Slovenia; department of Entomology, National Museum of Natural History, NHB-105, Smithsonian Institution, P.O. Box 37012, Washington, DC 20013-7012, USA; department of Zoology & Entomology, University of the Free State, P.O. Box 339, Bloemfontein 9300, South Africa; department of Biology, University of Ljubljana, P.O. Box 2995, SI- 1001 Ljubljana, Slovenia Abstract. Analysis of ecological data of the arboricolous nephilid spider Clitaetra irenae Kuntner 2006, endemic to Maputaland forests, South Africa, indicates the species’ dependence on this highly threatened habitat. We tested C. irenae habitat dependence via GIS analysis by plotting the known distribution against southern African ecoregions. In the southern part of its range, C. irenae inhabits almost exclusively one ecoregion. the Maputaland coastal plain forests; but further north, in tropical southern Africa, it continues inland into Malawi’s woodlands. We test and refute the hypotheses that C. irenae inhabits exclusively mature trees, trees of a particular species, trees with a smooth bark, tree habitats at certain height above ground, and only closed canopy forest stands. The ecological niche of C. irenae is flexible as long as suitable trees under at least partially closed canopy are available. We quantify the C. irenae ontogenetic web changes from orb to ladder and the simultaneous hub displacement towards the top frame. Such web allometry allows the web to increase vertically but not horizontally, which enables the spider to remain on the same tree throughout its development and thus the ladder web architecture is an adaptation to an arboricolous life style. Adult hub displacement, common in spiders with vertical webs, is explained by gravity. Clitaetra irenae web orientation on trees correlates with forest closure, and might indicate the Maputaland forest quality. We argue for utilization of the ecology of arboricolous nephilid orb-weaving spiders (Clitaetra and Herennia ) in systematic conservation assessments in the Old World tropics. Keywords: Behavioral ecology, evolution, conservation. Maputaland, South Africa, Herennia Opsomming. Analise van ekologiese data van die boombewonende nephilid spinnekop Clitaetra irenae Kuntner 2006, endemies tot Maputaland woude in Suid-Afrika, dui die spesie se afhanklikheid van hierdie hoogs bedreigde habitat aan. Ons het C. irenae habitat afhanklikheid via GIS analiese getoets deur die bekende verspreiding teen die suider-Afrikaanse ekologiese streke aan te teken. In die suidelike deel van die verspreiding kom C. irenae in slegs een ekostreek, die Maputaland kusvlakte woude, voor, maar verder noord in die verspreiding kom dit verder in die binneland, in Malawiese bosveld. voor. Ons toets en verwerp die hipotesisse dat C. irenae slegs volgroeide borne bewoon, borne van ‘n spesifieke spesie, borne met gladde bas, boomhabitatte op ‘n sekere hoogte van die grond af, en slegs toe kroonbedekking woude. Die ekologiese nis van C. irenae is aanpasbaar solank daar geskikte borne onder ten minste gedeeltelike toe kroonbedekking beskikbaar is. Ons kwantifiseer C. irenae ontogenetiese webveranderings van ‘n wawiel- na ‘n leerweb en die gelyktydige verskuiwing van die kern na die bokant van die raam. Sulke web allometrie gee die spinnekop die vermoe onr die web vertikaal te vergroot sender horisontale veranderings, wat die spinnekop in staat stel om op dieselfde boom te bly regdeur sy ontwikkeling. Ons sien die leerweb dus as ‘n aanpassing tot boomlewende gewoontes. Volwasse kern verskuiwing, algemeen in spinnekoppe met vertikale webbe, word deur gravitasie verduidelik. C. irenae se web orientasie op borne korreleer met kroonbedekking, en kan ‘n aanduider wees van Maputaland woud kwaliteit. Ons stel voor dat nephilid wawiel-web spinnekoppe (Clitaetra en Herennia) se ekologie gebruik word in sistematiese bewarings assesserings in die Oil Wereld tropiese gebiede. Identifying areas of high conservation value (systematic conservation assessment sensu Knight et al. 2006) is one of the main priorities of modern ecology and conservation biology. Traditionally, conservation assessments have been largely based on vertebrate biodiversity data, while data on more diverse groups such as arthropods remain poorly utilized. This is particularly true for species-rich tropical faunas, where the natural history and ecology of most arthropod species continue to be unstudied. Among spiders, a clade with nearly 40,000 described species (Platnick 2007) and many more projected (Coddington & Levi 1991; Coddington 2005), potential indicator species are rarely identified. Most recent studies have investigated the impacts of pollution, disturbance, 5 Corresponding author. E-mail: kuntner@gmail.com and habitat classification on spiders at the community level, while species-level studies have focused on effects on predation ecology and heavy metal assimilation (Marc et al. 1999). However, these studies primarily dealt with the European and American faunas, while examples from the Afrotropical region are unknown. Here, we investigate the ecology of the recently described Clitaetra irenae Kuntner 2006, an arboricolous nephilid spider endemic to southern African forests, and use the newly acquired behavioral and ecological data to test its dependence on a highly threatened habitat - the Maputaland coastal plain forests. Maputaland is an ecological-geographical entity comprising the coastal plain of north-eastern parts of KwaZulu-Natal (South Africa) from Richards Bay (28°48'S, 32°05'E) in the south to Xai-Xai in southern Mozambique (25°02'S, 34°25'E) 583 584 THE JOURNAL OF ARACHNOLOGY in the north, and extending inland along the Lebombo Mountain range into eastern Swaziland (Olson et al. 2001; Van Wyk & Smith 2001). Due to the high levels of plant and vertebrate endemism (Van Wyk 1994, 1996; Van Wyk & Smith 2001) and the patchy distribution of unique habitats such as sand forest, Maputaland is of prime conservation importance. Furthermore, the composition of certain faunal and floral assemblages of different sand forest patches are significantly different (e.g., Kirkwood & Midgley 1999; Matthews et al. 1999, 2001; Van Rensburg et al. 1999, 2000), which supports the conservation of this particular habitat and its accompa- nying biological diversity throughout the region. Maputaland appears among six out of nine global biodiversity conservation priority areas (Brooks et al. 2006), being recovered within the crisis ecoregions, the biodiversity hot spots, the endemic bird areas, the centers of plant diversity, and the global 200 ecoregions, as well as within a megadiverse country. South Africa (for details see Brooks et al. 2006, and references therein). The Old World tropical nephilid spider genus Clitaetra Simon 1889 is important in araneoid systematics because of its phylogenetic position as the sister taxon to all other nephilids (Kuntner 2005, 2006); these, in turn, appear to be the basal araneoid lineage (Kuntner et al. 2008). While the large-bodied nephilids ( Nephila , Nephilengys ) are well studied (Kuntner 2007) and recognized as ecologically important in most (sub)tropical ecosystems, Clitaetra ecology has largely been unknown (but, see Kuntner 2006). Of the six known Clitaetra species only C. irenae (Fig. 1) biology has been studied in some detail — Kuntner (2006) focused on its web-building behavior and general natural history. Clitaetra irenae was found in South Africa’s subtropical coastal dune and sand forests of northern KwaZulu-Natal, with an outlying museum record from Malawi (Kuntner 2006: fig. 26). As in the Australasian nephilid spider genus Herennia Thorell 1877 (see Kuntner 2005), Clitaetra species are obligate tree dwellers building orb-webs against tree bark (Figs. 1, 2), with a few centimeters clearance between the planar web and the substrate. As in Herennia (Kuntner 2005), Nephilengys Koch 1872 (Japyassu & Ades 1998; Kuntner 2007), and Nephila Leach 1815 (Bleher 2000), juvenile Clitaetra webs resemble standard araneoid orb-webs (Figs. 2D, E). However, the webs of adult female C. irenae are highly modified ladder webs (Figs. 2A, C), defined as two dimensional orb-webs with parallel side frames and zig-zag sticky lines substituted for spirals (Kuntner et al. 2008:fig. 23). The species’ apparent dependence on a particular habitat (Maputaland forests) and microhabitat (mature trees), as indicated by historical distribution records and the distribution of these particular forests, prompted our detailed investigation into its behavioral ecology emphasizing its ontogenetic web architecture shifts. We investigate C. irenae web allometry by quantifying the developmental modification of the orb web into a ladder, pose a behavioral explanation for such an allometric shift, and deduce its ecological implications. Through explicit hypothe- ses (below) we examined whether or not the web modification in C. irenae correlates with a particular habitat type, with climatic conditions and/or altitude, and whether the species depends on mature Maputaland forests for sustainable populations. A species that is a narrow endemic of southern Africa and 'a narrow habitat specialist (an obligate tree dweller) in a threatened coastal plain habitat is potentially at risk of habitat loss, and could be added to the list of species used as indicators of habitat quality. Specifically, we tested the following hypotheses: (1) Endemism: Kuntner (2006) suggested that C. irenae was endemic to northern KwaZulu-Natal. However, such an hypothesis was defined geographically, not ecologically, and failed to explain the single outlying record from Malawi (Kuntner 2006). We tested whether C. irenae is endemic to the wider Maputa- land (including Mozambique and Swaziland) by using precisely defined ecoregions (Olson et al. 2001). (2) Habitat preference : Most material examined by Kuntner (2006) was collected in sand forests along the Maputaland coastal plain. We investigated Kuntner’s (2006) expectation that C. irenae is confined to mature sand forests. During our survey we noted the habitat type and canopy closure. (3) Tree preference : To test the spiders’ dependence on a particular tree species we attempted to identify all trees where spiders were found. We predicted the spiders would prefer a particular tree species as vaguely suggested by Kuntner (2006). (4) Tree size: If the species was confined to mature forests, we predicted that larger spiders would prefer larger trees, and that web size would correlate with spider size. (5) Substrate (bark) structure: To test if the spiders utilize certain types of substrate, we classified bark into three categories (smooth, medium, or rough). We expected that trees with smooth bark would be more likely to host these spiders as the uneven surface of rough bark may hinder two-dimensional web construction. (6) Website orientation: The spider’s website orientation on a tree might differ between closed canopy forests and partially open canopy tree stands (woodland/ thicket) due to sunlight penetration. Assuming that the species is indeed native to closed canopy forests and that inhabiting other types of tree stands is an artifact of habitat fragmentation, we expected that we would find no web orientation preference in closed canopy stands, where continuous direct sunlight is an exception. In contrast, in more open tree stands where webs could potentially be exposed to direct sunlight, web orientation should be away from the sun (to the south). (7) Distance from ground: Finally, we tested for differ- ences in web height above ground among the spider instars (size classes). We predicted that spiders would preferentially occupy lower positions on a tree as they age because larger tree trunk circumference closer to the ground would be more suitable for larger webs. KUNTNER ET AL. ECOLOGY OF CL1TAETRA IRENAE 585 Figure 1. — Clitaetra irenae from South Africa (A-C, Fanies Island, 2001; D, Ndumo Game Reserve, 2006): A. Female holotype (see Kuntner 2006) at the hub of her web; B. Female feeding at the hub of her web; note the egg-cases above the hub camouflaged with prey remains; a male was present in her web (not shown); C. Paratype male (see Kuntner 2006) at the hub of his orb web, feeding on dipteran prey; D. Subadult female at web hub, feeding on cricket. Scale bars = ! .0 mm. (8) Web allometry: Clitaetra irenae web architecture shifts during spider ontogeny from orb to ladder web. Thus, we predicted that the ladder index and the hub displacement index (both defined below) would increase with spider size. METHODS Field study. — We visited six reserves (Fig. 3) in KwaZulu- Natal (South Africa) between 9-30 April 2006, the period when adult Clitaetra irenae are known to occur (Kuntner 2006). These reserves (Ndumo Game Reserve, Tembe Ele- 586 THE JOURNAL OF ARACHNOLOGY Figure 2. — Clitaetra irenae web architecture: A-C. Adult female webs; A. Frontal view, note ladder shape of web with parallel side frames (F), “sticky spiral’’ (SS) not spiralling, with silk enforced hub (H) displaced towards top frame; B. same web, lateral view; C. Female (arrow) in web, with first instar offspring in upper web; above hub is empty egg sac. D. Juvenile web (second instar); note typical orb-web architecture with bent side frames (F), round sticky spiral (SS), and upper radii as long as lower ones. E. Another second instar web. with parameters measured (a = web width, b = web height, c = top frame to hub). Photographs by M. Kuntner taken in Ndumo Game Reserve, South Africa, 2006; web measurements (cm) in A, B: a = 8, b = 29, c = 10.5; D: a = 5, b = 11, c = 5.5. phant Park, Kosi Bay Nature Reserve, Sodwana Bay National Park, Hluhluwe-Imfolozi National Park, and the Greater St. Lucia Wetlands Park) slightly surpass the previously known C. irenae geographic range, except for the Malawi datum (Kuntner 2006), and therefore test the species’ geographical limits in Maputaland. Within the reserves we searched for the easily recognizable species (females, see Fig. 1A; males, see Fig. 1C; for identification details see Kuntner 2006) at various localities, focusing on all available habitat types. Where more than a single C. irenae spider was found, we measured the following ecological and behavioral parameters on all webs: date of collection; locality; site; habitat; latitude and longitude; host tree species (if known); tree bark structure categorized as smooth (as exemplified by Celtis africana ), KUNTNER ET AL. — ECOLOGY OF CLITAETRA IRENAE 587 KOSI BAY ; Maputaland coastal forest mosaic KwaZulu-Cape coastal forest mosaic [ J| Southern Africa mangroves Zambezian and Mopane woodlands Drakensberg montane grasslands, woodlands and forests Maputaland-Pondoland bushland and thickets other ecoregions HLUHLUWE A ^ SODWANA BAY ST. LUCIA ☆ (^) not found • 1958 A 1991, 1998 O 2001 2006 25 50 Figure 3. — Clitaetm irenae, currently known distribution plotted against southern African ecoregions. The outlying 1958 Malawi record (see Kuntner 2006) falls into “Zambezian and Mopane woodlands," which stretches south adjacent to “Maputaland coastal forest,” the species' prime ecoregion. The 1991-2001 data are from Kuntner (2006), new Field records (2006) are from this study. medium (as in Balanites maughamii), or rough (as in Acacia nigrescens ); tree trunk circumference at the level of spider web hubs; canopy coverage categorized as closed, partially open, or open; stage or instar number; web width (Fig. 2E), web height (Fig. 2E), distance from web hub to top frame (Figs. 2D, E) and web hub height above ground; website orientation (8 directions: N, NE, E, SE, etc.), and possible further comments. The First six (autecological) parameters document locality, habitat, and test hypothesis 2 (habitat preference). The next four parameters, i.e., host tree species, bark structure, trunk circumference and canopy coverage, are microhabitat data and test hypotheses 2-5. Spiders were assigned to seven categorical size classes (stages) corresponding to ontogenetic instar numbers (size correlates with age and instar number). Instar numbers were estimated from relative spider size, falling into fairly discrete size classes, assuming seven post-egg sac instars during female ontogeny: adult females were scored as 7, newly hatched animals 1, and others fell in between up to 6 (penultimate). Instars 2-7 build their own webs, but first instar spiderlings remain in the mother’s web (Fig. 2C). No developmental study has been done on Clitaetra. The assumption of seven instars in C. irenae ontogeny follows from data on other nephilids; e.g., the giant females Nephila pilipes (Fabricius 1793) go through about 10 juvenile instars while the small males go through about 4, the number of molts depending on the season and food availability (Higgins 2002). Three web parameters, namely web width, web height, and distance from hub to top frame, quantified developmental shifts in webs (see below). The last two parameters, distance from hub to ground and website orientation, indicate microhabitat preference and test hypotheses 6 and 7. In order to increase visual contrast, webs were dusted with cornstarch for measurement and photography. All measure- ments were taken with a tape measure and are reported in centimeters, unless noted otherwise. Web orientation was taken with a compass. Web architecture abbreviations are: H = hub (Figs. 2A, B, D, E); F = frame (Figs. 2A. D); RA = 588 THE JOURNAL OF ARACHNOLOGY radius (Figs. 2A, D); SS = sticky spiral (Figs. 2A, D); a = web width (Fig. 2E); b = web height (Fig. 2E); c = upper web height = distance top frame to hub (Fig. 2E). Web allometry. — We derived two ratios quantifying web allometry. The ladder index is the relative web height, defined as the ratio of web height to web width, and quantifies the transition from orb to ladder web during ontogeny. Similar to web shape sensu Zschokke (1993), the ladder index, as used here, differs in values increasing with extreme architecture. Hub displacement is lower orb/total web height using the formula (b-c)/b, where b = web height and c = top to hub (Fig. 2E). Hub displacement index is similar to web asymme- try indices of Masters & Moffat (1983), Rhisiart & Vollrath (1994), and Kuntner et al. (2008), but its values increase rather than decrease with the hub being eccentric towards the top web frame. Statistical analyses. — We explored the relationships between the spider size (instar) and the following variables: 1) web height, 2) web width, 3) distance of web from the ground, 4) tree circumference, 5) ladder index, and 6) hub displacement index. Web parameters were plotted as box-plots (Tukey 1977). Due to the low number of webs in stages 4 and 5, individual points were plotted for these stages. Differences in medians between stages 2 and 7 were tested using the Mann- Whitney U- test. Freely available statistical environment R (R Development Core Team 2006) was used for plotting and for significance tests. Web orientation on trees (eight main directions) was interpreted through circular statistics using the Rayleigh test (Fisher 1993). GIS analysis. — In order to test hypothesis 1 (endemism), we analyzed all available C. irenae locality data (Kuntner 2006; this study) using GIS. Our analysis builds on the digital base map by Bletter et al. (2004), which was obtained from the New York Botanical Garden (http://www.nybg.org/bsci/digital_ maps/), with permission from the authors. The base map, which derives from the Environmental Systems Research Institute (ESRI) data sets ArcWorld®, ArcAtlas®,and Digital Chart of the World® (http://www.esri.com/), was built specifically for the Neotropics (Bletter et al. 2004), but also contains detailed data on the World’s terrestrial ecoregions (Olson et al. 2001), courtesy of the World Wildlife Fund (http://www.wwf.org/). Olson et al. (2001) recognize 867 terrestrial ecoregions worldwide, which is a much improved resolution compared with previous attempts to classify terrestrial biotas. The map delimits six ecoregions within the geographical limits relevant to this study (Fig. 3). The C. irenae locality data (1958; 1991-2001 data are from Kuntner (2006); new field records (2006) are from this study) were superimposed on the southern African part of the base map using ArcView® GIS. RESULTS Clitaetra irenae occurred in all reserves (Fig. 3) except Hluhluwe-Imfolozi National Park, which falls into the Drakensberg montane grasslands, woodlands, and forests ecoregion (Fig. 3). In Kosi Bay and Sodwana Bay the spiders were present but their abundances per trees were too low to be measured. In some reserves, e.g., Ndumo and Tembe, the species was abundant. The outlying record of a single C. irenae male from Malawi (1958 datum from Kuntner 2006) falls on the ecoregion “Zambezian and Mopane woodlands,” which stretches south adjacent to “Maputaland coastal forest” (Fig. 3), the species’ main ecoregion. Figure 4 summarizes the data obtained in Ndumo Game Reserve, Tembe Elephant Park and the two localities in the Greater St. Lucia Wetlands Park. The full data set with exact geographical coordinates for each habitat within the reserves is published online at www.nephilidae.com. In total, we investi- gated 166 spiders and their webs, the majority (N = 1 18, 71%) being second instars; the numbers of other instars investigated were much lower (Fig. 4A). Habitat preference characterizes the four forest types where the webs were measured (Fig. 4B). We characterized the habitat at Nyamiti Hide (Ndumo Game Reserve), where the majority of the aggregated second instar webs were found, as “riverine bush” (“deciduous orthophyll scrub with trees” of De Moor et al. (1977); “subtropical bush” of Haddad et al. (2006)), since the tree stands had a partially open canopy, with the terrain sloping gradually towards the Nyamiti Pan. Fig. 4C show that the most data (133 individuals) were taken in forest stands with only partially open canopy (riverine bush at Nyamiti Hide was all partially open habitat), and that all sand forest patches examined (Ndumo, Tembe, St. Lucia) were closed canopy (33 data points). There was little overlap of tree species (if known) between localities. However, sand forest trees harboring spider webs had a smooth and medium bark texture, while those in riverine bush had rough or medium bark (Fig. 4D). Figures 5A-F show the. relationships of interval and ratio data by instar, and report the results of the non-parametric comparisons (Mann-Whitney U test) of instar 2 (« = 118) and 7 (n — 18). Web height and web width increased significantly with spider size (Figs. 5A, B , P < 0.001). These two relationships support our assumption that spider web size correlates with age (instar). No significant differences were found in the web distance from the ground (Fig. 5C, P - 0.39) among the spiders of different size. Contrary to our prediction, larger spiders did not prefer larger trees (Fig. 5D, P = 0.15). However, as predicted, both measures of web allometry, i.e., the ladder index (Fig. 5E, P < 0.001) and the hub displacement index (Fig. 5F, P < 0.001), increased with spider size. Figure 6 plots web orientation frequencies in eight catego- ries (north, northeast, east, etc.) using circular statistics (Fisher 1993). Pooling all data (Fig. 6A) shows that C. irenae webs were randomly distributed on all sides of trees. However, the closed canopy data (Fig. 6B) reveal a significant (Rayleigh test, P < 0.05) preference for the northern side of trees, and the partially open canopy data (Fig. 6C) plots as significantly bimodal (Rayleigh test, P < 0.05), the spiders preferring southern and eastern faces of trees. Similarly, we found that the webs on smooth bark (only found in closed canopy sand forests), showed a significant preference towards the north. However, the webs on medium bark (closed and partially open forests), and on rough bark (only in partially open canopy forest), showed random distributions. The northern orienta- tion of webs under closed canopy is related to canopy closure, not bark type, as both subsets of webs, the smooth bark webs and medium bark webs showed a significant preference for the northern side of trees (Rayleigh test, P < 0.05 and P < 0.1, respectively). KUNTNER ET AL. — ECOLOGY OF CLITAETRA I RENA E 589 Figure 4. — Field data summary (category, n, percentage): A. Individuals by instar; B. Habitat; C. Canopy coverage; D. Bark structure. DISCUSSION The disproportionately high numbers of small juveniles (instar 2) measured compared to other instars (Fig. 4A) may be explained by the fact that second instars tend to aggregate around the mother’s web and are thus more easily located than instars 3-6. One of the alternative explanations, the higher mortality of larger juveniles, is contradicted by the fact that 18 females versus fewer 4-6th instar juveniles were collected (Fig. 4A). Seasonality seems the best explanation: as our study focused on the known adult C. irenae phenology, our April field work only sampled a part of the species’ life cycle. Distribution and phenology. — The ecoregion with the ma- jority of C. irenae records (Fig. 3) is the Maputaland coastal forest mosaic, but some records also fall into the adjacent ecoregions, including the Southern Africa mangroves and the KwaZulu-Cape coastal forest mosaic. The localities we visited that fall into these latter two ecoregions were no different with regards to forest structure and tree and bark microhabitat from the ones falling into Maputaland, and thus we view these records as continuous with “larger Maputaland.” At this resolution, the borders between these ecoregions are fairly arbitrary. It should be noted, however, that two of the three South African localities visited where C. irenae was not found are “Drakensberg montane grasslands, woodlands and forests” (Fig. 3), whose forest microhabitat structure is quite different from Maputaland. These habitats are further inland and at higher altitudes than C. irenae typically inhabits, perhaps indicating particular microclimatic, altitudinal and habitat preferences. We predict that in the south the species inhabits Maputaland coastal forests into Mozambique, but that further north in tropical southern Africa, it continues inland into the adjacent ecoregion (the Zambezi an and Mopane woodlands) as far north as Malawi. Although no specimen records currently exist from Mozambique and Swaziland, we predict that the species occurs there. Our data support the endemism hypothesis by 1) showing the continuity of the Maputaland forest mosaic ecoregion into Mozambique, and by 2) showing the ecoregion Zambezian and Mopane woodlands’ adjacency to Maputaland. Where found, the C. irenae adult abundances were greatest during the beginning of our study (11-13 April 2006) while towards the end (28-29 April 2006) no further adults were found. Habitat preference. — No C. irenae webs were measured in open canopy stands (Fig. 4C), which is not to be interpreted to mean that the spiders never occur there. In Maputaland C. irenae inhabits most tree habitats, including lone trees in semi- open canopy areas, and even synanthropic vertical surfaces, but preferentially occupies partially open and closed canopy stands. However, three of the four habitats where spiders and their webs were investigated in detail are forests (Fig. 4B), which is consistent with our assumption that forests and not other types of tree stands are the species’ prime habitat. Evidently, the species is not confined to sand forests, refuting our hypothesis 2. We also conclude that C. irenae does not exclusively inhabit closed canopy forests (Fig. 4C), and is much more common in partially open stands (see below). 590 THE JOURNAL OF ARACHNOLOGY A Web Height p < 0.001 E o JQ 0 5 B Web Width CM CO 'sf o 4 5 Instar p < 0.001 6 7 3 Tree Size E CD O C CD CD E 0 1 o 0 0 o o CO o o CM o o p= 0.148 6 7 Instar Instar E Ladder Index p < 0.001 F Hub Displacement p< 0.001 Figure 5. — Clitaetra irenae web parameters by spider size (instar number): A. Web height; B. Web width; C. Distance from ground; D. Tree size preference; E. Ladder index; F. Hub displacement. Differences between stages 2 and 7 were tested (Mann- Whitney U test), shown as P values. Many trees with C. irenae webs (n = 71) could not be identified to species, but evidently C. irenae do not associate with one or a few particular tree taxa, nor with trees of a particular bark structure, refuting the hypotheses 3 and 5. This is reflected in the diverse habitats in which the species was encountered, each of which contains unique plant assem- blages. Our predictions that larger spiders chose larger, matured trees, and lower portions of trees, ignored web allometry and were based on the assumption that spider web size increases with age. While the measured webs indeed increased with age (Figs. 5 A, B) the fact that a variety of tree trunk sizes were utilized by each instar (Fig. 5D) refutes our hypothesis 4. Similarly, spiders of a certain age showed no preference for website height above ground (Fig. 5C) refuting the hypothesis 7. This would indicate a fair degree of tolerance to microclimatic variation with web height above ground (varying between 15.5 cm and 337.0 cm), which are expected to differ substantially from ground to canopy height. Dispersal of second instars and subsequent web construction KUNTNER ET AL.— ECOLOGY OF CLITAETRA IRENAE B Closed Canopy Web Orientation S C Partially Open Canopy Web Orientation N Figure 6. — Clitaetra irenae web orientation using circular statistics (Fisher 1993): A. All data (no orientation preference detected); B. Closed canopy forest (significantly north); C. Partially open canopy forest (significantly bimodal, south or east). Light versus dark grey graph colors represent statistically insignificant versus significant (Rayleigh test, P < 0.05) orientation. appears to be in the immediate vicinity, and on the same tree, as the mother’s web. Web allometry. — The two indices quantifying the ontoge- netic web changes from orb to ladder and the simultaneous hub displacement towards the top frame (Figs. 5E, F) both increase with spider size, which supports the prediction of 591 hypothesis 8. Such web allometry explains the lack of correlation between spider age and tree size. The ladder index, which is the relative web height, increases significantly with age, which is consistent with our initial observations of modified adult webs {Fig. 2). To a growing spider and its web the limiting factor on a tree of a given size is the horizontal website availability (tree circumference). The observed onto- genetic web allometry allows the growing spider’s web to increase vertically, while at the same time the corresponding horizontal web increase is not required, allowing the spider to remain on the same tree. The smallest tree on which an adult female web was constructed (tree circumference 42.0 cm, web width 6.5 cm) would suggest that the curvature of the tree trunk may also play a role in the web site selection, making available points of attachment on the bottom and top of the web but eliminating the restriction of horizontal attachment points on the side of the web. We see the ladder web as an adaptation to an arboricolous life style because it eliminates the need of a potentially costly and dangerous walk from one (smaller) tree to another (larger) tree. Ants, which are abundant on trees where C. irenae occur, may be important predators (Kuntner 2006). However, ants are not known to invade webs. Thus, orb-web spiders may be safer from predators on (or close to) their webs than walking about. The webs of all adult nephilid spiders have displaced hubs, mostly towards the top (for rare horizontal displacement in Nephilengys , see Kuntner 2007). Hubs are often displaced above the web center in other araneoid spiders with vertical webs (see Kuntner et al. 2008), notably in araneids. The logical explanation for an ontogenetic shift from a central hub in small juveniles towards the eccentricity seen in larger, heavier spiders is gravity. Masters & Moffat (1983) demonstrated that predation success of the araneid spider Larinioides sclopetarius (Clerck 1757) improves in webs with hubs displaced above the web center, as the time to reach the prey upwards and downwards is thus optimized. Ladder webs on tree trunks, present in the extant species of the basal Cliiaetra-Herennia grade, were the ancestral life style of the pan tropical clade Nephilidae and reversed to aerial orb webs in the ancestor of Nephila (Kuntner et al. 2008). All nephilids, C. irenae included, retain the non-sticky spiral (NSS) in their finished web unlike most other orbweavers (Kuntner 2006; Kuntner et al. 2008). NSS or auxiliary spiral functions as a guide during the spider’s sticky spiral construction (Zschokke 1993). Unlike in Nephila and Nephi- lengys, NSS in Clitaetra webs is difficult to discern, perhaps because these spirals are thin and may get stuck with the narrowly meshed sticky ones. Kuntner (2006) determined the NSS presence by observing the spiders build at night (and not cutting the NSS when laying the sticky spiral), and the same could not be determined by photographs of finished webs. The retention of the NSS in nephilids has been suggested to have evolved in response to female gigantism (Hormiga et al. 1995). However, NSS seems to have been present in the nephilid ancestor, where sexual dimorphism was only moderate and not extreme (Kuntner et al. 2008). Perhaps the retention of NSS was originally related to ladder web architecture, and its presence in derived nephilids represents evolutionary time lag. The convergent presence of NSS in Scoloderus webs (below) may also be related to ladder web architecture. 592 THE JOURNAL OF ARACHNOLOGY While ladder web architecture is homologous in Clitaetra and Herennici (Kuntner et al. 2008), it is clearly not related to, and differs functionally from, other known araneoid ladder webs. For example, neotropical Scoloderus builds an extreme ladder and, as in nephilids, retains the NSS (Eberhard 1975: fig. 2), but this web is aerial, not arboricolous, and its hub is displaced to the lower, not upper frame of the web, making the spirals above the hub resemble a ladder. The hub of the extreme ladder web of an unidentified araneoid from New Guinea described by Robinson & Robinson (1972), however, is displaced up as in nephilids, but that web is aerial and apparently lacks the NSS. Furthermore, while nephilid ladder webs are permanent structures (Kuntner 2005, 2006; Kuntner et al. 2008), the above webs are taken down daily and rebuilt every night (Robinson & Robinson 1972; Eberhard 1975). Ladder webs apparently evolved convergently in araneids and nephilids, perhaps in order to exploit new websites (trees) or food resources. The first may be particularly true for nephilids: while no study has focused on the prey of Clitaetra and Herennia these spiders exploit tree trunks as websites and Herennia even evolved a unique web design using pseudoradii (Kuntner 2005; Kuntner et al. 2008). Eberhard (1975) pointed out that a ladder web architecture allows for a more constant mesh size than any circular orbweb, which may be related to specialization for certain prey types. Aerial ladder webs have been suggested to represent convergent adaptation for ensnaring moths (Eberhard 1975; Stowe 1978). However, prey capture in tropical spiders is often anecdotal. While Stowe (1978) showed that 68% of 212 prey items of Scoloderus were moths, Robinson & Robinson (1972) observed a single prey item, a moth, attracted into the ladder web of their unknown spider by a light source. Although the focus of our study was not a quantification of C. irenae prey, the few sporadically observed insect prey items in their webs (Fig. 1A-D: Orthoptera, Diptera, Lepidoptera (Pyralidae), Homoptera) suggest that the species is an opportunistic predator not particularly specialized on moths. Web orientation. — Superficially, C. irenae webs appear randomly oriented (Fig. 6A). However, preferential spider web orientation occurs in both closed and partially open canopy forests. In closed canopy forests (Fig. 6B) the spiders show a preference for the northern side of the tree, refuting the first part of the web orientation hypothesis (that no orientation preference would be shown in closed canopy forests). Two types of bark that were found under closed canopy (smooth, medium) both show a significant orientation to the north, which indicates that northern orientation of webs under closed canopy is related to canopy closure, not bark type. Conversely, in partially open canopy forests (Fig. 6C) the spiders preferentially chose southern and eastern faces of trees, which is supportive of the second part of our hypo- thesis 6. Web orientation probably affects web microclimate. Pro- longed direct sun exposure could harm the spider or its web, or may affect prey availability and/or predation pressure. In the southern hemisphere, the predicted preference for the south- ern, shady side of trees makes sense in a partially open forest. It is somewhat surprising, however, that the spiders inhabiting closed canopy forests seem to prefer a northern (perhaps warmer) orientation, where we predicted randomness, though this result does not, per se, refute hypothesis 6. Since we mainly scored smooth barked trees in sand forests (all closed canopy) it is not surprising that the web orientation patterns on smooth bark trees resemble those of closed canopy forests (Fig. 6B). The preference for the southern and eastern sides of trees in partially open canopy habitats (Fig. 6C) would suggest an aversion for direct sunlight and a preference for darker sides of trees. Implications for Maputaland ecology. — Previous Maputa- land studies have focused on organisms such as dung beetles and birds to assess community heterogeneity and the impacts of habitat destruction and regeneration on faunal and floral components (Van Rensburg et al. 1999, 2000; Davis et al. 2002; Wassenaar et al. 2005). However, Maputaland arachnids are highly diverse, with 457 species recorded from the Ndumo Game Reserve, 10.1 12-ha in size (Haddad et al. 2006). Given such diversity it is likely that several species are endemic to Maputaland and could be used as indicators of changes and disturbances to habitats, including sand forest. Spiders are generally more sensitive than other arthropod groups to the vegetative structural conditions in a habitat, particularly web-builders (e.g., Marc et al. 1999; Stiles & Coyle 2001; Finch 2005). In monitoring or habitat evaluation using spiders as bio-indicators, the absence of species typical for a habitat (stenotypic species) is often indicative of low habitat quality, including vegetation structure (Bonte & Maelfait 2001). Thus, the identification of a sensitive species in the Maputaland fauna could provide an indication of the current condition of forest patches and the extent of disturbance to which the local fauna is presently being exposed. Once identified, indicator spider species can also be used for long- term monitoring of landscapes (Bonte et al. 2002), which will enable conservationists to assess changes in the condition of these forests over time. A particularly important current conservation issue in South Africa is the impact of heavy utilization of sand forests by elephants in the Tembe Elephant Park, Maputaland. This is thought to lead to the irreversible opening up of the sand forest into a structure comparable to mixed woodland (W.S. Matthews pers. comm.), which is likely to have a strongly negative long-term effect on the plant and animal communities, and on the diversity of the sand forest. Additionally, the growing rural human population in Maputa- land is putting increasing pressure on sand forest patches outside conservancies (Kirkwood & Midgley 1999). In Africa, particularly, forests are opened up by big game, and such forest gaps are not necessarily indicative of unnatural disturbance. Indeed, in the reserve where elephants occur (Tembe), partially open canopy patches seem to be continuous with closed canopy sand forests. Thus, our finding of most individuals of C. irenae in partially open canopy forest stands does not refute our hypothesis about C. irenae dependence on the undisturbed Maputaland forest habitats. This could be an artifact of our under-sampling of closed canopy locations. Assuming that indigenous, undisturbed Maputaland forests are closed canopy, and thus the original preference of C. irenae spiders in a quality habitat is towards the north, scoring web orientation might be indicative of forest quality/disturbance. The easy C. irenae identification (Fig. 1, see Kuntner 2006) warrants further investigation as to whether it might be a suitable bioindicator. KUNTNER ET AL. ECOLOGY OF CLITAETRA IRENAE Conclusions. — The data at hand suggest the ecological and behavioral dependence of C. irenae on the threatened Maputaland forests. The wider Maputaland endemism hy- pothesis receives support, but the hypotheses that C. irenae inhabits exclusively sand forests, mature trees, trees of a particular species, trees with a smooth bark, tree habitats at certain height above ground, and only closed canopy forest stands, are refuted. Evidently the species’ ecological niche is flexible to an extent but requires suitable tree habitat under at least partially closed canopy. However, the web orientation on trees appears to be indicative of closed versus partially open canopy forest. Conservationists may benefit from utilizing the available arthropod data in assessing the quality of tropical forests. The ecology of obligate arboricolous orb-weaving spiders (like the nephilids Clitaetra and Herennia), seems especially well suited for systematic conservation assessments in the (sub)tropics because they range from western Africa (Kuntner 2006) through South and Southeast Asia into Australasia (Kuntner 2005) where some species are narrow endemics and others appear to be widespread and invasive (Kuntner 2005, 2006). The "Africa + Asia + Australasia” tropical belt matches the maps of global biodiversity conservation priorities (Brooks et al. 2006), but also lies precisely in a zone of high population pressure and low human development index (see Jha & Bawa 2006) , a detrimental combination of factors associated with heavy deforestation. ACKNOWLEDGMENTS We thank Jonathan Coddington, Ingi Agnarsson, Jeremy Miller, Wayne Matthews, Tatjana Celik, and Simona Kralj- Fiser for discussions and detailed comments on an early draft, Marjan Jarnjak for producing the species distribution map, Douglas C. Daly (New York Botanical Garden, Bronx) for kindly authorizing the use of the base map of Bletter et al. as our GIS source, and Ansie Dippenaar-Schoeman, Xander Combrink, Sharron Hughes, and Irena Kuntner for help, assistance, or advice. The comments of S. Toft, S. Zschokke and an anonymous reviewer further improved our paper. This research, made possible through a Ezemvelo KZN Wildlife collecting and export permit no. 1010/2006, was supported by the Slovenian Research Agency (grant Z 1-7082-06 18 to MK) and the EU 6th Framework Programme (Marie Curie grant MIRG-CT-2Q05-036536 to MK). LITERATURE CITED ap Rhisiart, A. & F. Vollrath. 1994. Design features of the orb web of the spider, Araneus diadematus. Behavioral Ecology 5:280-287. Bleher, B. 2000. Development of web-building and spinning apparatus in the early ontogeny of Nephihi madagascariensis (Vinson, 1863) (Araneae: Tetragnathidae). Bulletin of the British Arachnological Society 11:275-283. Bletter, N., J. Janovec, B. Brosi & D.C. Daly. 2004. A digital base map for studying the Neotropical Bora. Taxon 53:469-477. Bonte, D., L. Baert & J.-P. Maelfait. 2002. Spider assemblage structure and stability in a heterogeneous coastal dune system (Belgium). Journal of Arachnology 30:331-343. Bonte, D. & J.-P. Maelfait. 2001. Life history, habitat use and dispersal of a dune wolf spider (Pardosa monticola [Clerck, 1757] Lycosidae, Araneae) in the Flemish coastal dunes (Belgium). Belgian Journal of Zoology 131:145-157. 593 Brooks, T.M., R.A. Mittermeier, G.A.B. da Fonseca, J. Gerlach. M. Hoffmann, J.F. Lamoreux, C.G. Mittermeier, J.D. Pilgrim & A.S.L. Rodrigues. 2006. Global biodiversity conservation priori- ties. Science 313:58-61. Coddington, J.A. 2005. Phylogeny and classification of spiders. Pp. 18-24. In Spiders of North America: an Identification Manual. (D. Ubick, P. Paquin, P.E. Cushing & V. Roth, eds.). American Arachnological Society. Coddington, J.A. & H.W. Levi. 1991. Systematics and evolution of spiders (Araneae). Annual Review of Ecology and Systematics 22:565-592. Davis, A.L.V., R.J. Van Aarde, C.H. Scholtz & J.H. Delport. 2002. Increasing representation of localized dung beetles across a chronosequence of regenerating vegetation and natural dune forest in South Africa. Global Ecology and Biogeography 11:191-209. De Moor, P.P., E. Pooley, G. Neville & J. Barichievy. 1977. The vegetation of Ndumo Game Reserve, Natal: a quantitative physiognomic survey. Annals of the Natal Museum 23:239-272. Eberhard, W.G. 1975. The ‘inverted ladder’ orb web of Scoloderus sp. and the intermediate orb of Eustala (?) sp. Araneae: Araneidae. Journal of Natural History 9:93-106. Finch, O.-D. 2005. Evaluation of mature conifer plantations of secondary habitat for epigeic forest arthropods (Coleo- ptera: Carabidae; Araneae). Forest Ecology and Management 204: 21-34. Fisher, N.I. 1993. Statistical Analysis of Circular data. Cambridge University Press, Cambridge, UK. 298 pp. Haddad, C.R., A.S. Dippenaar-Schoeman & W. Wesolowska. 2006. A checklist of the non-acarine arachnids (Chelicerata: Arachnida) of the Ndumo Game Reserve, Maputaland, South Africa. Koedoe 49:1-22. Higgins, L.E. 2002. Female gigantism in a New Guinea population of the spider Nephila maculata. Oikos 99:377-385. Hormiga, G., W.G. Eberhard & J.A. Coddington. 1995. Web- construction behavior in Australian Phonognatha and the phylog- eny of nephiline and tetragnathid spiders (Araneae: Tetragnathi- dae). Australian Journal of Zoology 43:313-364. Japyassu, H.F. & C. Ades. 1998. From complete orb to semi-orb webs: developmental transitions in the web of Nephilengys cruentata (Araneae: Tetragnathidae). Behaviour 135:931-956. Jha, S. & K.S. Bawa. 2006. Population growth, human development, and deforestation in biodiversity hotspots. Conservation Biology 20:906-912. Kirkwood, D. & J.J. Midgley. 1999. The floristics of sand forest in northern KwaZulu-Natal, South Africa. Bothalia 29:293-304. Knight, A.T., A. Driver, R.M. Cowling, K. Maze, P.G. Desmet, A.T. Lombard, M. Rouget, M.A. Botha, A.F. Boshoff, J.G. Castley, P.S. Goodman, K. Mackinnon, S.M. Pierce, R. Sims-Castley, W.I. Stewart & A. Von Hase. 2006. Designing systematic conservation assessments that promote effective implementation: best practice from South Africa. Conservation Biology 20:739-750. Kuntner, M. 2005. A revision of Herennia (Araneae: Nephilidae: Nephilinae), the Australasian ‘coin spiders.' Invertebrate System- atics 19:391-436. Kuntner, M. 2006. Phylogenetic systematics of the Gondwanan nephilid spider lineage Clitaetrinae (Araneae, Nephilidae). Zoolo- gica Scripta 35:19-62. Kuntner, M. 2007. A monograph of Nephilengys, the pantropical ‘hermit spiders’ (Araneae, Nephilidae, Nephilinae). Systematic Entomology 32:95-135. Kuntner, M„ J.A. Coddington & G. Hormiga. 2008. Phylogeny of extant nephilid orb-weaving spiders (Araneae, Nephilidae): testing morphological and ethological homologies. Cladistics 24:147-217. Marc, P., A. Canard & F. Ysnel. 1999. Spiders (Araneae) useful for pest limitation and bioindication. Agriculture, Ecosystems and Environment 74:229-273. 594 THE JOURNAL OF ARACHNOLOGY Masters, W.M. & A.J.M. Moffat. 1983. A functional explanation of top-bottom asymmetry in vertical orbwebs. Animal Behaviour 31:1043-1046. Matthews, W.S., A.E. Van Wyk & N. Van Rooyen. 1999. Vegetation of the Sileza Nature Reserve and neighbouring areas, South Africa, and its importance in conserving the woody grasslands of the Maputaland Centre of Endemism. Bothalia 29:151-167. Matthews, W.S.. A.E. Van Wyk, N. Van Rooyen & G.A. Botha. 2001. Vegetation of the Ternbe Elephant Park, Maputaland, South Africa. South African Journal of Botany 67:573-594. Olson, D.M., E. Dinerstein, E.D. Wikramanayake, N.D. Burgess, G.V.N. Powell, E.C. Underwood. J.A. D’Amico, I. Itoua, H.E. Strand, J.C. Morrison, C.J. Loucks, T.F. Allnutt, T.H. Ricketts, Y. Kura, J.F. Lamoreux, W.W. Wettengel, P. Hedao & K.R. Kassem. 2001. Terrestrial ecoregions of the World: a new map of life on Earth. BioScience 51:933-938. Platnick, N.I. 2007. The World Spider Catalog, Version 8.0. American Museum of Natural History, New York. Online at http://research.amnh.org/entomology/spiders/catalog/. R Development Core Team. 2006. R: A language and environment for statistical computing. Version 2.4.0 R Foundation for Statistical Computing, Vienna, Austria. Online at http://www. R-project.org. Robinson, M.H. & B. Robinson. 1972. The structure, possible function and origin of the remarkable ladder-web built by a New- Guinea orbweb spider (Araneae: Araneidae). Journal of Natural History 6:687-694. Stiles, G.J. & F.A. Coyle. 2001. Habitat distribution and life history of species in the spider genera Theridion , Rugathoides , and Wamba in the Great Smoky Mountains National Park (Araneae, Theridiidae). Journal of Arachnology 29:396^112. Stowe, M.K. 1978. Observations of two nocturnal orbweavers that build specialized webs: Scoloderus cordatus and Wixia ectypa (Araneae: Araneidae). Journal of Arachnology 6:141-146. Tukey, J.W. 1977. Exploratory Data Analysis. Addison-Wesley Publishing, Reading, Massachusetts. 688 pp. Van Rensburg, B.J., S.L. Chown, A.S. Van Jaarsveld & M.A. McGeogh. 2000. Spatial variation and biogegraphy of sand forest avian assemblages in South Africa. Journal of Biogeography 27:1385-1401. Van Rensburg, B.J., M.A. McGeogh, S.L. Chown & A.S. Van Jaarsveld. 1999. Conservation of heterogeneity among dung beetles in the Maputaland Centre of Endemism, South Africa. Biological Conservation 88:145-153. Van Wyk, A.E. 1994. Maputaland-Pondoland region. Pp. 227-235. In Centres of Plant Biversity: A Guide and Strategy for Their Conservation. (S.D. Davis, V.H. Heywood & A.C. Hamilton, eds.). Oxford University Press. Oxford, UK. Van Wyk, A.E. 1996. Biodiversity of the Maputaland Centre. Pp. 198-207. In The Biodiversity in African Savannahs. (L.J.G. Van der Maesen, X.M. Van der Burgt & J.M. Van Medenbach de Rooy, eds.). Kluwer Academic Publishers, Dordrecht. Van Wyk. A.E. & G.F. Smith. 2001 . Regions of floristic endemism in Southern Africa: a review with emphasis on succulents. Urndaus Press, Hatfield, South Africa. 160 pp. Wassenaar, T.D., R.J. Van Aarde, S.L. Pimm & S.M. Ferreira. 2005. Community convergence in disturbed subtropical dune forests. Ecology 86:655-666. Zschokke, S. 1993. The influence of the auxiliary spiral on the capture spiral in Araneus diadematus Clerck (Araneidae). Bulletin of the British Arachnological Society 9:169-173. Manuscript received 14 July 2007, revised 28 January 2008. 2008. The Journal of Arachnology 36:595-599 Differential survival of Geolycosa xera archboldi and G. hubhelli (Araneae, Lycosidae) after fire in Florida scrub James E. Carrel1: Division of Biological Sciences, 209 Tucker Hall, University of Missouri-Columbia, Columbia, Missouri 6521 1-7400, USA. E-mail: carrelj@missouri.edu Abstract. A replicated pre- and post-burn study of survival of small and large Geolycosa xera archboldi McCrone 1963 and G. hubhelli Wallace 1942 in Florida scrub was conducted. These two syntopic species were chosen because G. x. archboldi prefers large gaps of barren sand in the scrub matrix, sites with little fuel for fires, whereas G. hubhelli strongly favors small gaps having some leaf litter, sites with modest or high fuel-loads. On the basis of these species-specific differences in microsite characteristics, I hypothesized that G. x. archboldi would be very fire tolerant but that G. hubbelli would be fire intolerant. I established two size classes for the Geolycosa : small spiders had 3-5 mm diameter X 5-9 cm deep burrows; large spiders had > 6 mm diameter X 10-17 cm deep burrows. Burrows of 25 spiders in each species X size class were marked before a burn in seven burn units ( = fire management areas) and survival or mortality of each occupant was ascertained over the course of 5 days post-burn. Thus, the experimental design was 2 species X 2 size classes X 7 burn units X 25 replicates/burn unit (n = 700 spiders total). Survivorship was very high in small and large G. x. archboldi and in large G. hubbelli (93-96%), but it was low in small G. hubbelli (35%). Temperature recordings suggest mortality in small G. hubbelli was caused by high temperatures at depths of 5-10 cm during intense, but brief burns that characterize fires in Florida scrub. In contrast, large G. hubbelli had burrows sufficiently deep so that most of them did not experience lethal temperatures during bums. Keywords: Burrowing wolf spiders, endemism. Lake Wales Ridge, body size, fire ecology Florida scrub is a fire-prone ecosystem confined to ancient sand ridges in the peninsular part of the state. This ecosystem also supports biotic communities that comprise a globally important, imperiled center of endemism (Deyr up 1989; Deyrup & Eisner 1993; Dobson et al. 1997; Menges 1999; Marshall et al. 2000; Estill & Cruzan 2001; Weekley et al. 2008). Presumably, as part of the suite of characters needed to survive in scrub, endemic species have evolved adaptations to frequent landscape-level burns that rapidly consume the leaf litter and standing vegetation. For example, the dominant woody shrubs have most of their biomass below ground, so they survive and quickly regenerate the shrub matrix by sprouting. In contrast, most endemic herbs are killed by fire and post-burn increases in abundance are due to seedling recruitment (Weekley & Menges 2003, and references therein). Scrub animals have three common methods for coping with fire at a landscape scale. On the one hand, some such as sand skinks ( Plestiodon reynoldsi), gopher tortoises ( Gopherus polyphemus), and flightless pygmy mole crickets ( Neotridacty - lus archboldi), persist in place by exploiting a subterranean life style in the sandy soils (Robbins & Myers 1992; Deyrup 2005). On the other hand, the Florida scrub jay ( Aphelocoma coerulescens) and other highly dispersive animals flee the oncoming flames on wing or foot and settle in unburned scrub (Robbins & Myers 1992). A third approach, one used by weak-flying insects and arboreal spiders, such as the red widow spider ( Latrodectus bishopi Kaston 1938), is to experience high mortality during a bum and to recolonize subsequently from nearby, unburned refugia (Deyrup & Eisner 1996; Carrel 2001, 2008). Two species of rare burrowing wolf spiders, Geolycosa xera archboldi McCrone 1963 and G. hubbelli Wallace 1942, are 'Current address: Archbold Biological Station, 123 Main Drive, Venus, Florida 33960, USA. endemic to oak scrub on the Lake Wales Ridge in the middle of peninsular Florida (Marshall et al. 2000). Because the spiders spend most of their lives below ground in tubular burrows they construct in the sand, I expected that they might be fire tolerant, similar to other subterranean animals. But knowing that small individuals build much shallower burrows than larger, older individuals (Table 1 and Figure 1), I hypothesized that survival of a burn in Geolycosa might be size dependent because smaller spiders build more shallow burrows than larger spiders and. as a result, small spiders could be more exposed to lethal temperatures that penetrate the upper layer of soil when scrub is burned. In addition, because G. x. archboldi prefers large (> 1 nr), barren gaps of sand and does not decorate its burrow entrance with a turret, whereas G. hubbelli favors small gaps (~ 0.1 nr) in the shrubby matrix having leaf litter from which it obligatorily builds a turret (Carrel 2003a, b), I also hypothesized that the latter species might be more likely to perish in a fire. To test these ideas, I conducted a pre- and post-fire study of survival (or mortality) of individual G. x. archboldi and G. hubbelli in two size classes (small and large individuals, Tables 1 and 2) over the course of several burn events in Florida scrub. I also collected ambient temperate data in Geolycosa burrows and on the soil surface during a fire. To my knowledge this is the first replicated, quantitative study of survivorship in any spider exposed to burning of its habitat, and it may be one of the few such studies with any terrestrial arthropod to date (Warren et al. 1987; Whelan 1995; Siemann et al. 1997; Swengel 2001). METHODS Study site. — I conducted a pre- and post-fire study of Geolycosa survival in flat, oak scrub at the 2101 ha Archbold Biological Station, in southern Highlands County, Florida (elev. 36-46 m, 27°H'N, 81°21'W). The work was performed in the most extensive vegetative association, called scrubby 595 596 THE JOURNAL OF ARACHNOLOGY Table 1.- -Depth and volume of burrows constructed by small and large Geolycosa spiders. Typical data were calculated using best-fit regression equations published by Carrel (2003a). Burrow size class (diameter, mm) Burrow Small Large Spider species properties (3-5) (6-15) G. xera archboldi Depth (cm) 4. 6-8. 2 10.0-15.4 Volume (cc) 0.4-2. 3 4.5-19.7 G. hubbelli Depth (cm) 5.6-9. 1 10.3-16.6 Volume (cc) 0.7-2. 8 4.5-49.0 flatwoods, which has fire-resistant slash pines (Pinus elliotti) scattered in a dense matrix dominated by low-growing shrubby oaks ( Quercus inopina , Q. chapmanii, and Q. geminate/), palmettos ( Serenoa repens and Saba I etonia, Arecaceae), and shrubby lyonias ( Lyonia ferruginea , L. fruticosa, and L. lucida, Ericaceae) (Abrahamson et al. 1984). For management purposes, the scrub at Archbold is organized into a series of 187 burn units and a detailed history of burning in each unit is available (Main & Menges 1997; unpublished Archbold records). 1 was able to work in seven units, ranging in size from 4.6 to 66.5 ha, two of which were burned in February 2001, one in October 2002, two in July 2007, and two in August 2007. Voucher specimens of both Geolycosa species were deposited in the Invertebrate Collec- tion at Archbold. Experimental design. —I haphazardly located 25 small (3- 5 mm diam.) and 25 large (6-15 mm diam.) burrows of both Geolycosa xera archboldi Geolycosa hubbelli Figure 1. Silhouettes of small and large burrows of Geolycosa xera archboldi and G. Iiubbelli prepared from representative plaster casts (Carrel 2003a). Note interspecific difference in the architecture of burrow bases. Table 2. — Attributes of two Geolycosa species placed into two size classes (small and large) based on diameter of their burrow openings, for study of survivorship after fire. Typical data were calculated using best-fit regression equations (Carrel 2003a). Sample size in this study («) for each species X size class is also given. Spider size class G. x. archboldi (turret absent) G. hubbelli (turret present) Small Burrow diameter (mm) 3-5 3-5 Carapace width (mm) 1. 4-2.3 1. 5-2.1 Body mass (mg) 8-30 8-17 Sample size (n) 175 175 Large Burrow diameter (mm) 6-10 6-15 Carapace width (mm) 2. 7-4.3 2.4-5. 1 Body mass (mg) 40-230 25-600 Sample size (n) 175 175 Geolycosa species by visually searching in seven different burn units 1-2 days before each was burned. Burrows were > 10 m from the perimeter of a burn unit to avoid edge effects, particularly kerosene-induced flames from drip torches used to ignite the leaf litter and vegetation. In previous studies (Carrel 2003a) I showed that the persistently open, circular burrows render these spiders very detectable: by conducting a rapid, but thorough visual search of an area (10-100 nr), one typically locates 90-95% of individuals actually present. Furthermore, the presence or absence of a turret constructed from leaves and debris, held in place with silk around the burrow opening, is a reliable tool for telling the species apart (Carrel 2003a). In addition, burrow diameter, as measured with calipers to the nearest 0.1 mm, is a highly reliable surrogate for the size of the occupying spider as well as the depth and volume of its burrow (Tables 1 and 2). Before a burn, I marked the location of each spider burrow (n = 700 total) by placing two thin metal stakes vertically in the sand —10 cm on opposite sides of the burrow entrance. Following a burn, I revisited each burrow for 5 consecutive days and determined if the resident spider was alive. I used four criteria for survivorship: sighting of a spider sitting near the top of its burrow; luring a spider from the burrow by the presence of insect prey that I tethered on a thread near the entrance; restoration of a damaged burrow entrance or turret; and placement of newly excavated sand on the ground near a burrow. If all these criteria were negative, on the fifth or sixth day post-fire I carefully excavated a spider’s burrow looking for its body. In so doing I could confirm that the burrow was occupied by a spider and, based on the soft, decomposing condition of a corpse, that the resident individual perished during or shortly after the blaze. Air and soil temperature measurements.- I used Hobo U-12 digital dataloggers (Onset Computer Corporation, Pocasset, Massachusetts) fitted with Type K thermocouples to record air and soil temperatures in the scrub, following the established protocols of Wally et al. (2006). After calibrating each machine, I programmed the dataloggers in the laboratory to record one reading per second and to output maximum temperatures at 1 min intervals prior to deployment in the scrub. I obtained two sets of temperature data: maximum daily temperatures inside G. x. archboldi burrows and nearby CARREL— SURVIVAL OL GEOLYCOSA AFTER FIRE 597 in undisturbed soil on hot, sunny days; and the intensity and duration of fire at point sites on the soil surface in oak scrub in order to gain a better perspective of the thermal dynamics experienced by subterranean spiders. The first set of temperature recordings was designed to determine whether the open burrows of small and large G. x. archboldi under typical summer daytime conditions were significantly warmer than intact soil at comparable depths in the scrub. I chose to study only G. x. archboldi because this species occurs predominantly in large, barren gaps of unshaded sand where solar heating is the most intense in scrub. In contrast, G. hubbelli is typically found in small gaps with leaf litter on the sand, so its burrows are insulated from solar heating both by the leaf litter and by shade cast by the surrounding shrub matrix. Thus, my reasoning was that if maximum daytime temperatures in open G. x. archboldi burrows were comparable to those in undisturbed soil at comparable depths, then a similar burrow/soil equivalency probably would hold for G. hubbelli (even though the maxima obviously would be smaller). (Subsequent measurements showed this relationship was valid, JEC unpubl. data.) Over the course of 3 weeks in late August-early September 2007 I simultaneously set up ten replicate sets for 1 day each with thermocouples in five different positions: at 0, 5 cm, and 10 cm depth in intact sand and at the bottom of small (3- 5 mm diam. X 3. 5-5. 2 cm depth) and large (6-12 mm diam. X 10.5-14.3 cm depth) G. x. archboldi burrows after the resident spiders were removed. Maximum daily air temperatures at 1.5 m above ground were also obtained at the official Archbold weather station on the days that soil temperatures were recorded. Secondly, in an attempt to characterize the intensity and duration of a fire in oak scrub, I acquired data on soil surface temperatures during a “category 3” burn in August 2007 from the plant ecology group at Archbold. (“Category 3”, the highest intensity in the classification scheme used by Archbold staff, means that most surface litter was consumed, all leaves of palmettos and shrubs 0-2 m elevation were completely consumed, and small twigs on shrubs were consumed in a blaze.) Following their published protocol (Wally et al. 2006), many thermocouples attached to datalogggers were placed in contact with the soil surface at a variety of locations to record soil surface temperatures during a burn event. Using data from ten dataloggers in sites that experienced heavy burns. I normalized the temporal records so that the peak maximum temperatures all occurred at the 10 min mark, so that there would be several min of pre-burn data as well as > 30 min post-maximum peak data. By definition, ignition threshold is > 60° C and cessation of fire is set at < 60° C; the 60° C benchmark is used because it corresponds to the temperature at which plant cell death occurs (Wally et al. 2006, and references therein). Statistical analyses. — I used the General Linear Models program of SPSS to perform ANOVA to evaluate the significance of variables in the sets of data on spider survival (SPSS 2005). The Levene test statistic was calculated to confirm that the variance did not differ significantly between the groups (P > 0.05). Differences in post-burn survivorship of spiders were analyzed by Chi square tests with Yates correction for small sample size (X2C, Simpson et al. 1960). Table 3. — Effect of burn event, species identity and body size of spiders (as measured by burrow diameter) on post-burn survival of two Geolycosa species in Florida scrub. Source of variation df MS F P Burn event 6 0.178 1.966 0.068 Species 1 14.573 160.9 < 0.001 Size of spider 1 18.241 201.4 < 0.001 Species X size 1 15.156 167.3 < 0.001 Error 672 0.091 I calculated the average (mean ± SE, n — 10) and range of maximum daily temperatures at all five locations in soil and in the air. 1 used the General Linear Models program of SPSS to perform univariate ANOVA to evaluate the significance of location in data on soil temperatures. The Levene test statistic was calculated to confirm that the variance did not differ significantly between the groups ( P > 0.05). Subsequently I performed two post hoc multiple range tests (Student-New- man-Keuls (SNK) and Tukey HSD) to determine in a pair- wise fashion which locations had significantly different temperatures (P set at 0.05) (SPSS 2005). After normalizing the soil surface temperature data during one burn event so that temperatures peaked at all locations {n = 10) in the 10th minute, I calculated the minimum, mean, and maximum temperature minute by minute for 30 min. RESULTS Post-burn survival of Geolycosa species. — Spider species, spider size, and spider species X spider size interaction were all highly significant variables determining the post-burn survival of Geolycosa species (Table 3). This meant that there was a complex interaction between spider species identity and spider size that required further analysis. Fortunately, as there were no significant differences among the seven burns according to the AVOVA results (Table 3), I was able to combine the data and delete “burn event” as a variable, which greatly simplified further analyses. As shown in Table 4, few small G. hubbelli (35.4%) survived the burns. In contrast, I found almost all large G. hubbelli (94.5%) and almost all G. xera archboldi regardless of size (small = 93.1%, large = 96.0%) were alive 5 days post-burn in the scrub. The intraspecific, size-depen- dent difference in survivorship for G. hubbelli was highly significant (X2C = 133.49, df — 1, P < 0.0001). Maximum daily temperatures in G. x. archboldi burrows. On ten sunny days in late summer 2007, maximum air temperatures at the Archbold weather station were hot, averaging 34.6 ± 0.3° C (mean ± SE, range 33.3-36.1 C), Table 4. — Survivorship of Geolycosa spiders as a function of burrow/body size and species identification. Results of statistical analyses (Chi square test with Yates correction for small sample size, X2C) for intraspecific size-based differences in survival are given. % Surviving Species Small (» = 175) Large (n = 175) X2 c P G. xera archboldi 93.1 96.0 0.89 NS G. hubbelli 35.4 94.5 133.49 < 0.0001 598 THE JOURNAL OF ARACHNOLOGY Time (min) Figure 2.- -Intensity and duration of a fire at a point source in Florida scrub. Note rapid onset and rise to peak temperature at soil surface, followed by somewhat less rapid decline. See methods for details. but the maximum daily temperature on the surface of fully exposed sand in scrub was much greater: nearly 16° C hotter (51.2 ± 0.9, 47.8-56.4 C). Univariate ANOVA showed there was a significant difference among the soil temperature data by location (F445 = 3.599, P = 0.013). Post hoc analysis revealed that, despite intense solar heating, the maximum daily temperatures at the bottoms of spider burrows and down in undisturbed soil remained significantly lower than at the surface (SNK and Tukey HSD tests, P < 0.05). Small, shallow spider burrows got as warm as soil at 5 cm depth (burrow: 38.5 ± 0.4, 36.7-40.3° C; soil: 37.8 ± 0.5, 35.7- 40.0° C; P = 0.44). In addition, large, deep spider burrows stayed even cooler (P < 0.05) than shallow ones during the day and their maximum daily temperatures were the equivalent to those in soil at 10 cm depth (burrow: 33.3 ± 0.4, 31.0-36.0° C; soil: 34.1 ± 0.4, 32.7-36.0° C; P = 0.28). Hence, despite the fact that the spiders’ burrows remained constantly open, the most extreme thermal climate experi- enced by resident animals if they were deep in the burrows would be virtually the same as if they were buried in undisturbed soil at a comparable depth, far less than that at the burrow entrance. Soil surface temperature during a burn. The time course of a burn in the scrub at any point in the burn unit was remarkably rapid. As shown in Fig. 2, the fire went from ignition temperature (60° C by definition) to peak maximum soil temperature (609-846 C) in < 2 min, then it declined to ~ 60 C in another 7 min. Hence, from the perspective of a Geolycosa hiding in its burrow, the fire lasted < 10 min. DISCUSSION Mortality in G. hubbelli. The results were generally in agreement with my initial hypotheses with one exception: the post-burn survivorship of large G. hubbelli was much greater than expected. In fact, to my surprise, it matched that for small or large G. x. archboldi (93-96%). 1 suspect burrow architecture makes large G. hubbelli very fire tolerant. As G. hubbelli grow they construct burrows that are not only wider in diameter and deeper, but also they excavate increasingly large, ovoid chambers at the bottoms (Table 1 and Fig. 1). Such bulbous refugia > 10 cm below the surface evidently protect large G. hubbelli from the brief but intense fires in the leaf litter and shrubbery above them, probably because the intense heat fails to penetrate to this depth. I think the cause of mortality in small G. hubbelli is not fire-induced asphyxiation. Under natural conditions in sandy Florida soils, extensive measurements of prevailing gases in — 16 cm deep burrows occupied by a closely related burrowing wolf spider, G. micanopy Wallace 1942, showed no significant increases in C02 or decreases in 02 concen- trations relative to ambient atmospheric values (Anderson & Ultsch 1997). Thus, during a fire in Florida scrub, I doubt there would be extensive depletion of oxygen down in the spiders' porous burrows. Moreover, detailed physiological studies by Prestwich (1983a, b; 1988a, b) have demonstrated that active Florida spiders rely mostly on anaerobic metabolism because nearly all of their tissue phosphagen is quickly (within 10-15 s) depleted after onset of activity. Hence, a 10-min period of hypoxia during a fire in Florida scrub should, at best, present Geolycosa spiders at rest in their burrows only with a mild respiratory challenge. I suspect the primary cause of fire-induced mortality in small G. hubbelli is high temperature in surficial soils and burrows. Field measurements show that soil temperatures at 2-3 cm depth rise to 80° C during intense fires in scrub, and at depths —5 cm the temperature may reach 65° C when fuel-loads are modest (< 0.6 kg dry leaves and stems on the ground/m2) (Hierro & Menges 2002; Alexis et al. 2007). However, if the fuel-load on the ground in Florida scrub is high (~1 kg/m"), as often is the case near burrows of G. hubbelli , then maximum soil temperatures at 5 cm depth during a burn are very hot (88 ± 9° C) (Hierro & Menges 2002). Several other studies have reported similar relation- ships between fuel load and soil temperature profiles (Whelan 1995). Hence, the relatively shallow burrows of small G. hubbelli probably reach temperatures that exceed the upper lethal temperatures of spiders, which range from 45 to 55° C for most species (Pulz 1987; Hanna & Cobb 2007). Assessment of fire effects on Geolycosa populations. — The strengths of this study are: 1. burn events were true replicates spanning 7 months of the calendar year; 2. pre- and post-burn sampling of many (u = 700) individual spiders was conducted; 3. sampling was size-based and quantitative. These attributes set it apart from almost all other previous studies that suffer from no replication or pseudoreplication and from nonquantitative or semiquanti- tative sampling methods (Warren et al. 1987; Siemann et al. 1997; Swengei 2001; van Mantgem et al. 2001; Hanula & Wade 2003). However, as explicitly pointed out by Whelan (1995), this study did not involve censuses of burrowing wolf spider populations before and after fires at randomly chosen sites. Thus, I cannot make any conclusions about whether fire has a significant impact on Geolycosa popula- tions in Florida scrub. But the data in this study suggest fire probably is not at all deleterious to populations of G. x. archboldi and it may have only a weak negative effect in the short-term on G. hubbelli populations. Long-term studies still in progress will address this subject (JEC, unpubl. data). CARREL-SURVIVAL OF GEOLYCOSA AFTER FIRE 599 ACKNOWLEDGMENTS I thank the Archbold Biological Station and its staff, especially H. Swain, M. Deyrup, E. Menges, C. Weekley, F. Lohrer, and K. Main, for providing long-term technical, financial, and intellectual support of the highest caliber. Additional support came from a Development Gift Fund at the University of Missouri. Finally, I deeply appreciate the continuing encouragement for extended field studies provided by Jan Weaver and other members of my family. LITERATURE CITED Abrahamson, W.G., A.F. Johnson, J.N. Layne & P.A. Peroni. 1984. Vegetation of the Archbold Biological Station, Florida: an example of the southern Lake Wales Ridge. Florida Scientist 47:209-250. Alexis, M.A., D.P. Rasse, C. Rumpel, G. Bardoux, N. Pechot, P. Schmalzer, B. Drake & A. Mariotti. 2007. Fire impact on C and N losses and charcoal production in a scrub oak ecosystem. Biogeochemistry 82:201-216. Anderson, J.F. & G.R. Ultsch. 1987. Respiratory gas concentrations in the microhabitats of some Florida arthropods. Comparative Biochemistry and Physiology 88A:585-588. Carrel, J.E. 2001. Population dynamics of the red widow spider (Araneae: Theridiidae). Florida Entomologist 84:385-390. Carrel, J.E. 2003a. Ecology of two burrowing wolf spiders (Araneae: Lycosidae) syntopic in Florida scrub: burrow/body size relation- ships and habitat preferences. Journal of the Kansas Entomolog- ical Society 76:16-30. Carrel, J.E. 2003b. Burrowing wolf spiders, Geolycosa spp. (Araneae: Lycosidae): gap specialists in fire-maintained Florida scrub. Journal of the Kansas Entomological Society 76:557- 566. Carrel, J.E. 2008. The effect of season of fire on density of female garden orbweavers (Araneae: Araneidae: Argiope) in Florida scrub. Florida Entomologist 91:332-334. Deyrup, M.A. 1989. Arthropods endemic to Florida scrub. Florida Scientist 52:254-270. Deyrup, M. 2005. A new species of flightless pygmy mole cricket from a Florida sand ridge (Orthoptera: Tridactylidae). Florida Ento- mologist 88:141-145. Deyrup, M. & T. Eisner. 1993. Last stand in the sand. Natural History 102(1 2):42-47. Deyrup, M. & T. Eisner. 1996. Description and natural history of a new pygmy mole cricket from relict xeric uplands of Florida (Orthoptera: Tridactylidae). Memoirs of the Entomological Society of Washington 17:59-67. Dobson, A.P., J.P. Rodriguez, W.M. Roberts & D.S. Wilcove. 1997. Geographic distribution of endangered species in the United States. Science 275:550-553. Estill, J.C. & M.B. Cruzan. 2001. Phytogeography of rare plant species endemic to the southeastern United States. Castanea 66:3-23. Hanna, C.J. & V.A. Cobb. 2007. Critical thermal maximum of the green lynx spider, Peucetia viridans (Araneae, Oxyopidae). Journal of Arachnology 35:193-196. Hanula, J.L. & D.D. Wade. 2003. Influence of long-term dormant- season burning and fire exclusion on ground-dwelling arthropod populations in longleaf pine flatwoods ecosystems. Forest Ecology and Management 175:163-184. Hierro, J.L. & E.S. Menges. 2002. Fire intensity and shrub regeneration in palmetto-dominated flatwoods of central Florida. Florida Scientist 65:51-61. Main, K.N. & E.S. Menges. 1997. Archbold Biological Station fire management plan. Land Management Publication 97-1. 104 pp. Marshall, S.D., W.R. Hoeh & M.A. Deyrup. 2000. Biogeography and conservation biology of Florida’s Geolycosa wolf spiders: threat- ened species in endangered ecosystems. Journal of Insect Conser- vation 4:1 1-21. Menges, E.S. 1999. Ecology and conservation of Florida scrub. Pp. 7-22. In Savannas, Barrens and Rock Outcrop Plant Communities of North America. (R.A. Anderson, J.S. Fralish & J.M. Baskin, eds.). Cambridge University Press, New York. Prestwich, K.N. 1983a. Anaerobic metabolism in spiders. Physiolog- ical Zoology 56:1 12-121. Prestwich, K.N. 1983b. The roles of aerobic and anaerobic metabolism in active spiders. Physiological Zoology 56:122-132. Prestwich, K.N. 1988a. The constraints on maximal activity in spiders. I. Evidence against the fluid insufficiency hypothesis. Journal of Comparative Physiology B 158:437-447. Prestwich, K.N. 1988b. The constraints on maximal activity in spiders. II. Limitations imposed by phosphagen depletion and anaerobic metabolism. Journal of Comparative Physiology B 158:449-456. Pulz, R. 1987. Thermal and water relations. Pp. 26-55. In Ecophysiology of Spiders. (W. Nentwig, ed.). Springer-Verlag, Berlin. Robbins, L.E. & R.L. Myers. 1992. Seasonal effects of prescribed burning in Florida: a review. Tall Timbers Research Inc., Tallahassee, Florida. Miscellaneous Publication Number 8. 96 pp. Siemann, E., J. Haarstad & D. Tilman. 1997. Short-term and long- term effects of burning on oak savanna arthropods. American Midland Naturalist 137:349-361. Simpson, G.G., A. Roe & R.C. Lewontin. 1960. Quantitative Zoology. Revised edition. Harcourt, Brace & World. New York. 440 pp. SPSS. 2005. SPSS for Macintosh, Release 11.0.4. SPSS Incorporated, Chicago, Illinois. Swengel, A.B. 2001. A literature review of insect response to fire, compared to other conservation managements of open habitat. Biodiversity and Conservation 10:1 141-1 169. van Mantgem, P., M. Schwartz & M.-B. Keifer. 2001. Monitoring fire effects for managed burns and wildfires: coming to terms with pseudoreplication. Natural Areas Journal 21:266-273. Wally, A.L., E.S. Menges & C.W. Weekley. 2006. Comparison of three devices for estimating fire temperatures in ecological studies. Applied Vegetation Science 9:97-108. Warren, S.D., C.J. Scifres & P.D. Teel. 1987. Response of grassland arthropods to burning: a review. Agriculture, Ecosystems and Environment 19:105-130. Weekley, C.W. & E.S. Menges. 2003. Species and vegetation responses to prescribed fire in a long-unburned, endemic-rich Lake Wales Ridge scrub. Journal of the Torrey Botanical Society 130:265-282. Weekley, C.W., E.S. Menges & R.L. Pickert. 2008. An ecological map of Florida’s Lake Wales Ridge: a new boundary delineation and an. assessment of post-Columbian habitat loss. Florida Scientist 71:45-64. Whelan, R.J. 1995. The Ecology of Fire. Cambridge University Press, Cambridge, UK. 346 pp. Manuscript received 22 January 2008, revised 2 May 2008. 2008. The Journal of Arachnology 36:600 BOOK REVIEW Harvestmen: the Biology of Opiliones. Edited by Ricardo Pinto-da-Rocha, Glauco Machado and Gonzalo Giribet. 2007. Harvard University Press, Cambridge, Massachusetts. 597 pp. ISBN 1 3:978-0- 674-02343-7. US$125. Except for some of the very small orders, nearly all arachnids have now received at least a first book on “The Biology of...” It is a curious question why it took so long for such a book to appear on the third most diverse order, Opiliones. But here it is at last, and it proves to be well worth the wait. Twenty-five authors have contributed to 15 chapters which summarize virtually everything that is known about harvestmen up to 2006. The organization of the book follows the general pattern: there are chapters on morphology, phylogeny and biogeography, systematics and paleontology, ecology, feeding, enemies and defense, reproduction, develop- ment, and social behavior. Each of the chapters is meticu- lously researched and, rather than simply recounting what is in the literature, the authors have synthesized and analyzed what they found. The result is that each chapter is an original review article that in itself is a significant contribution. Henceforward it will not be possible to write about or research harvestmen without referring to this book. And interestingly, the preponderance of South Americans among the chapter authors signals an important shift: the center of research on this group of arthropods has moved south, perhaps propelled by the enormous diversity in the group to be found in tropical America. A number of the chapters were of particular interest to this reviewer, especially the one on systematics by Pinto-da-Rocha and Giribet (it is also worth noting that one or the other of the editors has co-authored nine of the fifteen chapters in the book). This chapter is extraordinarily complete. Not only are keys to taxa included, but each described family is discussed in detail and abundantly illustrated. Keys are useful especially in the suborder Laniatores, where much reshuffling of families has taken place in the last decade. Reference is frequently made in these discussions to advances in our understanding of harvestmen systematics using phylogenetic data based on new molecular evidence. Areas requiring attention, such as the possibly paraphyletic family Ceratolasmatidae, are clearly pointed out. An overview of the current state of classification is given in a four-page table, covering the subfamily level, which also gives the numbers of genera and species currently included in each. Readers with long memories will recall that Ernst Mayr, in a text on taxonomy, used Opiliones as an example of an “over-split” group with, on average, less than two species per genus. The problem still exists in some places; the subfamily Tricommatinae has 51 species in 29 genera! Only one small quibble with this chapter — some of the many scanning electron micrographs used for illustrations are printed too small. The chapter on defense mechanisms is another gem, especially the section on chemical defenses, a signature feature of harvestmen biology. Here again, a useful chart puts in one place all the molecules and the species that produce them (except for the 22 gonyleptids studied by Hara et al. [2005]), and a quick perusal of that chart points the way for future work. Why, for example, does the single phalangiid studied so far (Phalangium opilio L. 1761) produce naphthoquinones, while the supposedly closely related sclerosomatine Leiobunum species produce long-chain alcohols and ketones? Research in the ecological chemistry of harvestmen has so far been focused on gonyleptomorph Laniatores, all of which produce a mixture of benzoquinones (at least 37 different molecules), while the chemistry of the defensive secretion is not known for even a single member of the Dyspnoi and is known for only one travunioid, Sclerobunus nondimorphicus Briggs 1971. Clearly this is an area of research where discoveries are waiting to be made, and one which I am currently exploring with a chemist colleague. Finally, it was fun to read the table on pp. 2-3, which lists vernacular names for harvestmen from more than 30 countries. Predominant are names that refer to harvest time, the long legs of the most obvious members of the order, and their perceived similarity to spiders. We also learn that it is only in Finnish in which the name for a species of Opiliones, lukki, means exactly that. This is an important and excellent book which should be in every arachnologist’s library, and which will be indispensable for university and departmental libraries. LITERATURE CITED Hara, M.R., A.J. Cavalhiero, P. Gnaspini & D.Y.A.C. Santos. 2005. A comparative analysis of the chemical nature of defensive secretions of Gonyleptidae (Arachnida: Opiliones: Laniatores). Biochemical Systematics and Ecology 33:1210-1225. William A. Shear: Biology Department, Hampden-Sydney College, Hampden-Sydney, Virginia 23943 USA. E-mail: wshear@hsc.edu 600 2008. The Journal of Arachnology 36:601-603 SHORT COMMUNICATION A new species of Xysticus (Araneae, Thomisidae) from Alberta, Canada Charles D. Dondale: Biodiversity, Research Branch, Agriculture & Agri-Food Canada, 960 Carling Avenue, Ottawa, Ontario K1A 0C6, Canada. E-mail: cjdondale@horizontech.ca Abstract. A new species of the crab-spider genus Xysticus (Thomisidae), X. albertensis, is described from northern Alberta, Canada. Specimens are compared with those of three species that closely resemble them and live in the same geographical region, namely, X. chippewa Gertsch 1953, X. canadensis Gertsch 1934, and X. britcheri Gertsch 1934. Keywords: Taxonomy, crab spider, western Canada The crab-spider genus Xysticus C.L. Koch 1835 was established for the widespread Palearctic species X. audax (Schrank 1803), and is currently regarded as a major world entity comprising more than 300 species (Platnick 2007). North America is home to nearly 70 species, a possible six of which are regarded as Holarctic (Dondale 2005; Dondale et al. 2006). The purpose of the present contribution is to describe a new species from northern Alberta. Canada, individuals of which closely resemble those of three species that are found in the same region, namely, X. chippewa Gertsch 1953, X. canadensis Gertsch 1934. and X. britcheri Gertsch 1934. Specimens are lodged in the following institutions: Canadian National Collection of Insects & Arachnids, Ottawa, Ontario, Canada (CNC); Strickland Entomological Museum, University of Alberta, Edmonton, Alberta, Canada (SEM). Figures 1-4. — Xysticus albertensis new species: 1. Palpus of holotype, ventral view; 2. Tegular apophyses of same, retrolateral view; 3. Epigynum of paratype, ventral view; 4. Spermathecae and copulatory tubes of same, dorsal view. ATS , atrial sclerite; BTA , basal tegular apophysis; CT, copulatory tube; DTA, distal tegular apophysis; E , embolus; RTA , retrolateral tibial apophysis; VTA, ventral tibial apophysis. Scale bar for Figures 1, 3, 4 = 0.20 mm, for Figure 2 = 0.08 mm. 601 Figures 5-13. — 5-7. Tegular apophyses of male palpi, retrolateral view: 5. Xysticus chippewa Gertsch; 6. X. canadensis Gertsch ; 7. X. britcheri Gertsch. 8-10. Atrial sclerites of female epigyna, ventral view: 8. X. chippewa; 9. X. canadensis; 10. X. britcheri. 11-13. Spermathecae and copulatory tubes, dorsal view: 1 1 . X. chippewa; 12. X. canadensis; 13. X. britcheri. CT , copulatory tubes. Scale bar for Figures 5-7 = 0.08 mm, for Figures 8-10 = 0.04 mm, for Figures 1 1-13 = 0.20 mm. Family Thomisidae Sundevall 1833 Genus Xysticus C.L. Koch 1835 Type species. — Aranea audax Schrank, 1803, original designation Xysticus albert ensis new species Figs. 1-4 “ Xysticus sp. 1 ”: Nordstrom & Buckle 2004:9. Type specimens. — Holotype male, paratype male, and paratype female from the margin of a small lake unofficially named “Esker Lake” by the collector, situated between Colin Lake (59°34'N, 1 1 0 08 ' W ) and Woodman Lake, in Colin-Cornwall Wildland Park, Alberta, Canada, 6-9 July 2002, Ted Johnson (CNC). Paratypes: 1 male, 1 female, with same data (CNC); 2 males, with same data (SEM). Etymology. -The name of the new species is derived from that of the Canadian province in which the type-series was collected. Diagnosis. — Individuals of X. albertensis new species closely resemble those of X. chippewa. but also bear some resemblance to those of X. canadensis and X. britcheri. the last three of which are currently treated as Holarctic. All four are characterized by the possession of a smoothly curved distal tegular apophysis that is neither angulate nor “heeled” (Figs. 1, 2, 5-7). Males of X. albertensis are distinguished from those of the other three species by the basally stout and abruptly hooked distal tegular apophysis (compare Fig. 2 with Figs. 5-7). Also, the two tegular apophyses in male X. albertensis are narrowly separated, whereas these structures in the other three species are more separate. Females of the four aforementioned species are characterized by possession of posteri- orly converging atrial sclerites (Figs. 3, 8-10). Individuals of X. albertensis differ from those of X. chippewa and X. canadensis by the possession of slender copulatory tubes (compare Fig. 4 with Figs. 11. 12) and from those of X. britcheri by thicker copulatory tubes (compare Fig. 4 with Fig. 13). The atrial sclerites of X. albertensis are only moderately separated posteriorly (Fig. 3), whereas those of female X. canadensis are well separated (Fig. 9), and those of X. britcheri are touching (Fig. 10). Individuals of both sexes of X. albertensis are predominantly dark brown (much as in X. britcheri), whereas those of X. chippewa and of X. canadensis are much paler. Description. — Holotype male: (Figs. 1, 2). Total length 3.98 mm; carapace 2.49 mm long, 2.16 mm wide. Carapace brown on yellowish background. Legs similar to carapace in color, paler distally; femur I with 8 erect macrosetae on prolateral surface. Sternum with many small round reddish brown conjoined spots. Abdomen dorsally with extensive dark brown areas; venter pale, with scattered reddish spots. Palpal tibia with stout tapered retrolateral apophysis and thick angular ventral apophysis; embolus moderately thick basally, slender distally, with free part separated from tegulum by distinct broad space distal to tegular apophyses; distal tegular apophysis thick at base, abruptly curved at tip, tapered to sharp point; basal tegular apophysis short, curved. Paratype female: (Figs. 3, 4). Total length 5.31 mm; carapace 2.41 mm long, 2.16 mm wide. Color much as in male, but somewhat paler, with eye area and median area of carapace creamy white; abdomen with only a few reddish or dark brown spots. Leg macrosetae as in male. Epigynum with shallow atrium; atrial sclerites converging posteriorly, moderately separated at posterior end (Fig. 3). Copulatory tubes approximately two-thirds as long as spermathecae; spermathecae shallowly lobed (Fig. 4). DONDALE— A NEW SPECIES OF XYSTICUS 603 Variation. — Males, n = 3: Total length 3.82-4.32 mm; carapace 2.49-2.57 mm long, 1.99-2.22 mm wide. Distribution. — Known only from the type locality. ACKNOWLEDGMENTS Donald J. Buckle first recognized Xysticus albertensis as new to science and kindly passed the specimens to the author for description. LITERATURE CITED Dondale, C.D. 2005. Thomisidae. Pp. 246-247. In Spiders of North America: an identification manual. (D. Ubick, P. Paquin, P.E. Cushing & V. Roth, eds.). American Arachnological Society. Dondale, C.D., T. Kronestedt & D.J. Buckle. 2006. Confirmation of the presence of Xysticus chippewa in Europe (Araneae, Thomisidae). Bulletin of the British Arachnological Society 13:361-369. Nordstrom, W. & D. Buckle. 2004. Spider records from Colin- Cornwall Lakes Provincial Park. Alberta Ministry of Community Development, Edmonton, Alberta, Canada. 31 pp. Platnick, N.I. 2007. The World Spider Catalog, Version 8.0. American Museum of Natural History, New York. Online at http://research.amnh.org/entomology/spiders/catalog/index. html. Manuscript received 7 February 2007, revised 12 February 2008. 2008. The Journal of Arachnology 36:604-605 SHORT COMMUNICATION An easy method for handling the genus Phoneutria (Araneae, Ctenidae) for venom extraction Leandro F. Garcia: Departamento de Biologia, Faculdade de Filosofia, Ciencias e Letras de Ribeirao Preto - Universidade de Sao Paulo, USP, Ribeirao Preto, Sao Paulo, Brazil. CEP 14040-900, Brazil. E-mail: lfgarc@gmail.com Luiz Henrique A. Pedrosa: Departamento de Bioquimica, Faculdade de Medicina de Ribeirao Preto - Universidade de Sao Paulo, USP, Ribeirao Preto, Sao Paulo, Brazil. CEP 14040-900, Brazil Denise R. B. Rosada: Departamento de Biologia, Centro Universitario Claretiano. Batatais, Sao Paulo, Brazil. CEP 14300-000, Brazil Abstract. This paper describes an easy, cheap, and safe method of capturing and handling the medically important spider Phoneutria for venom extraction. The method does not injure or kill the spider and allows the extraction of pure venom. Keywords: Armed spider, venomous spiders, vivarium The venomous ctenid spiders of the genus Phoneutria (P. nigriventer Keyserling 1891) (Figure la) are medically important due to their aggressiveness (Fig. la), their great speed, and the dangerous effects of their venom (Bticherl 1953a). Consequently, researchers are interested in obtaining pure venom from these spiders so that the medical significance of its components can be explored. Traditionally, spiders have been immobilized using anesthesia (CGj, chloroform, ether, or other substances) or with cold temperatures, but these techniques can cause mortality or alter the behavior and physiology of the animal (Harris et al. 1965; Randall 1982), with immediate or latent harmful effects (Nicolas & Sillans 1989). Furthermore, venom extraction following these handling methods has often entailed the excision and maceration of the entire venom gland, a procedure that produces impure venom (Biicherl 1953a). Biicherl (1953a) describes a technique that avoids these problems. It consists of irritating the spider so that it attacks and envenomates a device (two pipettes connected with an elastic surgical tube) from which the venom can be extracted. This may be the safest method for the extractor and does not cause the animal’s death, but the process is laborious and does not produce a sufficient amount of venom. Here we propose an easy, cheap method for capturing and handling Phoneutria for venom extraction that is safe for both the extractor and the spider. The handling device consists of a transparent or semi- transparent 2-liter plastic (PET) bottle (i.e., empty soft drink bottle) that has been cut transversely across the middle. Only the top half of the bottle is retained for use as a handling chamber (Fig. lb). An 8- cm longitudinal slit is cut into the side of the handling chamber. Then, while holding onto the top of the bottle mouth (with lid), the chamber is placed over the spider, imprisoning it. At this point, the spider typically attempts to climb upward toward the neck of the bottle. Filter paper should be used as a floor to capture any venom released by the spider; the venom can be recovered later by washing the filter paper with an organic solvent (e.g., acetonitrile). The spider is forced away from the bottle wall or lid and onto the filter paper by gently tapping the chamber against the extraction bench. Once the spider is on the filter paper, a glass stirring rod is inserted through the longitudinal slit and pushed down onto the spider in the area between the cephalothorax and the abdomen. This presses the spider onto the filter paper and immobilizes it (Figure lc). While continuing to press the animal down with the glass rod, the handling chamber is removed and the sides of the animal’s cephalothorax are grasped between the index finger and thumb (Figure Id). Once the spider is firmly grasped, the glass rod can be removed and the animal carried away for the venom extraction (Figure le). Venom extraction consists of placing electrodes against the region of the cephalothorax lying above the venom glands and then stimulating the spider with 6 V (Biicherl 1953b). The venom is collected from the chelicera in a small capillary tube (Smith & Micks 1968; Morris & Russell 1975) or on a glass plate. LITERATURE CITED Biicherl. W. 1953a. Novo processo de obtengao de veneno seco, puro, de Phoneutria nigriventer (Keyserling, 1891) e titulagao da LD50 em camundongos. Memorias do Instituto Butantan 25:153-174. Biicherl, W. 1953b. Escorpioes e escorpionismo no Brasil. I. Manutengao dos escorpioes em viveiros e extragao de veneno. Memorias do Instituto Butantan 25:53-82. Harris. R.L., R.A. Hoffman & E.D. Frazar. 1965. Chilling vs. other methods of immobilizing Hies. Journal of Economic Entomology 58:379-380. Morris, J.J. & R.L. Russell. 1975. The venom of the brown recluse spider Loxosceles reclusa : composition, properties and an im- proved method of procurement. Federation Proceedings 34:225. Nicolas, G. & D. Sillans. 1989. Immediate and latent effects of carbon dioxide on insects. Annual Review of Entomology 34:97-116. Randall, J.B. 1982. Surgical restraint apparatus for living spiders. Journal of Arachnology 10:91. Smith, C.W. & D.W. Micks. 1968. A comparative study of the venom and other compounds of three species of Loxosceles. American Journal of Tropical Medicine and Hygiene 17:651-656. Manuscript received 28 October 2007, revised 17 April 2008. 604 GARCIA ET AL. — METHOD FOR HANDLING PHONEUTRIA 605 Figure I . -The step-by-step handling method for venom extraction of Phoneutria nigriventer. a. The spider in an aggressive position; b. Left side of an empty 2-liter plastic bottle and right side of a bottle with a transverse cut and a glass stick inserted in the perpendicular cut; c. The trapped animal in the bottle being immobilized with the glass stick; d. Grasping the spider with fingers; e. Spider ready for venom extraction. 2008. The Journal of Arachnology 36:606-608 SHORT COMMUNICATION Courtship behavior and copulation in Tengella radiata (Araneae, Tengellidae) Gilbert Barrantes: Escuela de Biologi'a, Ciudad Universitaria Rodrigo Facio, Universidad de Costa Rica, San Pedro, San Jose, Costa Rica. E-mail: gilbert.barrantes@gmail.com Abstract. The first description of the courtship behavior and copulation is provided for Tengella radiata (Kulczynski 1909). The male courts the female by rocking his body and vibrating his abdomen. These behaviors seem to induce the female to move out from her retreat onto the sheet and incline her body to facilitate intromission. The female has an active role during the courtship: strumming the tunnel and sheet threads, apparently inducing the male to increase the frequency and intensity of his courtship. Palpal insertion is extremely short. The female terminates the copulation by lunging at the male. Keywords: Courting, sexual selection, funnel web spider Tengella radiata (Kulczynski 1909) has a wide distribution in Costa Rica where it inhabits mature and secondary wet forests and coffee plantations from 50 to 1500 m elev.( Wolff 1977; Santana et al. 1990, pers. obs.), but it is, nevertheless, unknown outside of this small country. Its web consists of a large, horizontal sheet with an upper tangle that contains some cribellate threads and a tunnel at the “interior” section of the sheet (Santana et al. 1990; Eberhard et al. 1993; Eberhard & Pereira 1993). Spiders rest near the tunnel opening during the day. The sexual biology of this spider is completely unknown. In nature males occasionally co-inhabit webs with adult females (W.G. Eberhard pers. comm), and I have occasionally observed males near or on the sheet of possibly adult females. Here I describe for the first time the courtship behavior and copulation of T. radiata and compare these behaviors to those of some species within families of the Tengella'' s sister groups lycosoids and agelenoids (Coddington 2005). The family is of interest because it is a cribellate member of the Lycosoidea. Courtship behavior and copulation of two pairs of T. radiata were filmed using a digital video camera Sony DCR-VX 1000 (30 frames/s). Both females were virgins raised from eggs in captivity and maintained in plastic boxes (30 X 18 X 1 1 cm) where they constructed their webs. One female was paired with one male that was also raised in captivity from different parents. The second female was paired with an adult male collected in the field. Male pre-copulatory and copulatory courtship behavior and copulation are defined as in Eberhard & Huber (1998). Male courtship refers to those behaviors that induce the female to respond in a way that favors the male's reproduction (Eberhard 1996). Copulation consists of all genitalic contact between a particular male-female pair, including the insertion of the embolus into the epigynal opening. It finishes when the pair separates from the copulatory position. Drawings were traced from video images. Spiders and egg sacs were collected near San Jose, Costa Rica; voucher specimens were deposited at the Museo de Zoologia, Escuela de Biologia, Universidad de Costa Rica. The first reaction of the male when placed on the female’s web was to walk on the sheet, more or less randomly at first and then toward the tunnel opening where the female rested. Following this movement, courtship and copulation can be roughly divided into three consecutive phases seen in both pairs: courtship by male, while the female is in the tunnel, with the result that the female moves out of the web tunnel; male courtship once the female is out on the sheet, presumably to induce the female to adopt the copulatory position; and copulation. The female responded to male courtship by either sending vibratory signals through the web threads, launching an apparent attack toward him, or adopting the copulatory position (described below). In total, the courtship behavior and copulations of T. radiata lasted 57 min in one pair (5 copulations) and nearly 90 min (9 copulations) in the second pair. In both cases the female eventually expelled the male from the web. When the male walked directly toward the tunnel opening, he stopped suddenly as the female began to strum the threads of the sheet with more or less alternate movements of her palps (Fig. 1). Strumming occurred in bouts of up to 20. During a strumming movement the female first extended both palps anteriorly and then flexed first one and then the other posteriorly, snagging some threads of the sheet with the tip of the palps. These sheet threads were pulled upward (visible in some video records) until they snapped free, possibly due to the tension. The palpal movements gave the impression of scratching the sheet surface rather than twanging particular threads. The female strumming movements apparently induced the male to stop, at least momentarily. The male then responded by rocking his body vigorously in antero-posterior direction (Fig. 2). He stood with all his legs on the sheet and shook the sheet visibly as he moved (occasionally legs I were lifted while rocking). The rocking movement was produced primarily by the antero-posterior movement of the male’s body rather than by bending his legs and occurred in bouts of 3-5 rocking movements ( n = 12). The male also frequently vibrated his abdomen once or twice just before a bout of rocking movements (5 out of 8 bouts in which illumination and angle were favorable). The female frequently stopped strumming the sheet (5 out of 7 sequences where both male and female were in focus); nearly immediately the male began to rock. If the female remained motionless after a rocking bout, the male often advanced a few millimeters toward her (2 of 8 instances). However, in most cases the female began again to strum the sheet as the male approached her, inducing him to stop and resume courtship. It seemed that in these cases the male rocked his body more vigorously before moving again toward the female that remained inside but near the tunnel opening. In one instance the female moved slowly out of the tunnel and stopped and strummed the sheet before continuing toward the male. The male rocked his body and advanced toward the female but she moved back a few millimeters and then darted at him in an attacking position (first legs slightly raised and directed forward and chelicerae spread). The male moved rapidly backward nearly 4 cm then began to rock his body again while the female returned to the tunnel. In both pairs, after several separate bouts of rocking by the male (12 in one pair and 17 in the other), the female moved out of the tunnel, ceased strumming the sheet, and allowed the male to approach 606 BARRANTES— SEXUAL BIOLOGY OF TENGELLA RADIATA Figure 1. — Strumming movements of the female’s palps on the sheet during courtship. Arrows show the sequence of the palps’ movements: dots- initial position, dashed- subsequent position, solid- final position. her. The male moved over the female’s body so that they faced opposite directions, and as he did so she inclined her body laterally (Fig. 3), with her epigynum exposed (copulatory position). From this position the male repeatedly contacted the female’s epigynum with one palp (Figs. 3, 4), while his other palp was held in front of his body. The male’s extended palp moved rapidly to touch the epigynum briefly and then withdrew (mean duration of extend-contact- withdraw cycle = 0.11 ± 0.02 s , n = 37). These movements apparently correspond to “flubs” observed in other species (Watson 1991; Stratton et al. 1996; Huber 1998; Eberhard & Huber 1998). On two occasions where the complete sequences were observed, the male made 17 and 29 flubs before the palp engaged with the epigynum and the haematodocha was finally inflated. The insertions with haema- todocha expansion lasted 0.38 s (± 0.16, n = 5). The haematodocha remained inflated during the entire insertion (n = 2). On three occasions (2 in one pair and 1 in the other, where the angle and focus were appropriate), the male was observed to move his abdomen up and down in possible copulatory courtship movements. These movements were slower than the pre-copulatory courtship vibratory Figure 2. — Two partial sequences (a and b) of male-female courtship. Black boxes- female strumming movements. Empty boxes- male rocking movements. Dashed boxes- male moving toward the female. 607 Figures 3, 4. — Position of male and female previous to and during insertion. 3. The female lays on her right side as the male walks over her body. 4. Male inserting his left pedipalp (embolus) in the left opening of the female’s epigynum. movements of the male’s abdomen, and were similar to the abdomen bobbing movement of Leucauge (Eberhard & Huber 1998). The pre- copulatory and copulatory courtship behaviors of a male after subsequent copulations on opposite sides were similar (e.g., flubs: 19- 15-16; three subsequent copulations). The extremely short insertions were successful as one of the females produced fertile eggs. The insertions seemed to be ipsilateral (by the position of the female’s epigynum and male’s palp), although 1 could not be completely certain due to the dark color of the epigynum and male’s palp. The female terminated copulation when she began to move her legs to stand on the sheet after a single successful insertion of the male’s palp (n = 5) or after several unsuccessful insertion attempts of the male’s palp (n = 4) (I could not differentiate unsuccessful insertion attempts from flubs). In one case the male remained over the female and she darted toward him in an attacking position. The female’s attack provoked an extremely rapid backward movement by the male that positioned him at 3 or 4 cm from her. After the female had ended the copulation, the male began a new approach with a sequence of pre-copulatory courtship behaviors. The courting male stopped frequently to pass his palps and sometimes his legs through his mouthparts before approaching her again; occasionally the male rubbed his palps against the sheet after grooming his palps with his chelicerae. During each new male approach, the female inclined her body toward the opposite side as the male walked over her and he immediately began to contact her epigynum with his other palp. Successive inclinations and insertions were on opposite sides (n = 12) except when the male failed to insert his palp in the previous attempt. In such cases the next male approach occurred on the same side of the female side that he had approached previously. The female was apparently responsible for the alternation of sides in subsequent copulations; it was clear on two occasions that she began to incline her body before the male could contact her. Having copulated, the female did not necessarily accept the male the next time he approached her. On several occasions (10 in one pair and 4 in the other) the male stopped his approach as she began to strum the sheet, and he restarted his rocking courtship behavior. Neither male charged his palps with sperm during the courtship, indicating that males charged their palps before encountering a female. The male’s pre-copulatory rocking and abdominal bobbing movements during copulation may reduce the female’s aggression and induce her to cooperate and use his sperm to fertilize her eggs (Eberhard 1996; Stratton et al. 1996; Eberhard & Huber 1998; Peretti et al. 2006). It is possible that these movements inform the female of the male’s quality. For example, one of the females of this study immediately lunged at and expelled from the web a small adult male (ca. 15% shorter than the males studied) that 1 had previously placed on her web. 608 THE JOURNAL OF ARACHNOLOGY The strumming behavior of the female, her ability to expel males with her attacks, and assumption of a distinctive acceptance posture, all clearly show her active role in mating (Peretti et al. 2006). This behavior possibly serves the female as a criterion for male selection as it induces the male to restart, and in some cases, seemingly to intensify his courtship behavior after he detects a female strumming. As in many other spiders, there was no indication of males being able to force females to cooperate (Huber 1996; Eberhard & Huber 1998). It is possible to compare some aspects of the courtship behavior of T. radiata with that of species of other related families: Agelenidae, Lycosidae and Pisauridae (Coddington 2005). In all these families, including T. radiata , males mount females facing in the opposite direction (Nielsen 1932; Miller & Miller 1987; Hebets et al. 1996; Stratton et al. 1996; Huber 1998; but see Bruce & Carico 1988). However in T. radiata the male’s courtship induces the female to incline her body to expose her epigynum, and thus his ventral surface touches (or nearly so) the ventral surface of the female, while in wolf spiders the male’s ventral surface touches the dorsal surface of the female and the male’s pedipalp reaches the female’s epigynum around the side of her abdomen. In at least some lycosids (Stratton et al. 1996) and in several agelenids (Huber 1998) males also flub (“scrape” in Stratton et al. 1996) repeatedly prior to insertions. Copulations involve ipsilateral palpal insertion on alternating sides in all families (Stratton et al. 1996). Alternating insertions with only a single expansion of the hematodocha also occurred in Rabidosa spp. (Lycosinae) (Stratton et al. 1996). Female attack behavior similar to that of T. radiata occurs in one lycosid (Miller & Miller 1987) and one pisaurid (Arnqvist 1992). Many of these behaviors are possibly homologous with those of Tengellidae, but information about many more species is required to make stronger arguments regarding the evolution of courtship behavior in Tengella and related families. Particularly, the information on courtship behavior of agelenoids, the sister group of Tengellidae and lycosoids, is very important to trace the evolution of the courtship behavior in these groups of spiders. I thank William G. Eberhard, Gail Stratton and two anonymous reviewers for many helpful comments on the manuscript and the Universidad de Costa Rica for financial support. LITERATURE CITED Arnqvist, G. 1992. Courtship behavior and sexual cannibalism in the semi-aquatic fishing spider Dolomedes fimbriatus (Clerck) (Ara- neae: Pisauridae). Journal of Arachnology 20:222-226. Bruce, J.A. & J.E. Carico. 1988. Silk use during mating in Pisaurina mira (Walkenaer) (Araneae, Pisauridae). Journal of Arachnology 16:1-4. Coddington, J.A. 2005. Phylogeny and classification of spiders. Pp. 18-24. la Spiders of North America: an identification manual. (D. Ubick, P. Paquin, P.E. Cushing & V. Roth, eds.). American Arachnoiogical Society. Eberhard, W.G. 1996. Female Control: Sexual Selection by Cryptic Female Choice. Princeton University Press, Princeton, New Jersey. 501 pp. Eberhard, W.G. & B.A. Huber. 1998. Courtship, copulation, and sperm transfer in Leucauge mariana (Araneae, Tetragnathidae) with implications for higher classification. Journal of Arachnology 26:342-368. Eberhard, W.G. & F. Pereira. 1993. Ultrastructure of cribellate silk of nine species in eight families and possible taxonomic implications (Araneae: Amaurobiidae, Deinopidae, Desidae, Dictynidae, Filis- tatidae, Hypochilidae, Stiphidiidae, Tengellidae). Journal of Arachnology 21:161-174. Eberhard, W.G., N.I. Platnick & R.T. Schuh. 1993. Natural history and systematics of arthropod symbionts (Araneae; Hemiptera; Diptera) inhabiting webs of the spider Tengella radiata (Araneae, Tengellidae). American Museum Novitates No. 3065:1-17. Hebets, E.A., G.E. Stratton & G.L. Miller. 1996. Habitat and courtship behavior of the wolf spider Schizocosa retrorsa (Banks) (Araneae, Lycosidae). Journal of Arachnology 24:141-147. Huber, B.A. 1998. Spider reproductive behaviour: a review of Gerhardt’s work from 1911-1933, with implications for sexual selection. Bulletin of the British Arachnoiogical Society 11:8-91. Miller, G.L. & P.R. Miller. 1987. Life cycle and courtship behavior of the borrowing wolf spider Geolycosa turricula (Treat) (Lycosidae). Journal of Arachnology 15:385-394. Nielsen, E. 1932. The Biology of Spiders: With Especial Reference to the Danish Fauna. Volume 1. Levin & Munksgaard, Copenhagen. 248 pp. Peretti, A., W.G. Eberhard & R.D. Briceno. 2006. Copulatory dialogue: females sing during copulation to influence male genitalic movements. Animal Behaviour 72:413-421. Santana, M., W.G. Eberhard, G. Bassey, K.N. Prestwich & R.D. Briceno. 1990. Low predation rates in the field by the tropical spider Tengella radiata (Araneae: Tengellidae). Biotropica 22:305- 309. Stratton, G.E., E.A. Hebets, P.R. Miller & G.L. Miller. 1996. Pattern and duration of copulation in wolf spiders (Araneae, Lycosidae). Journal of Arachnology 24:186-200. Watson, P.J. 1991. Multiple paternity as genetic bet-hedging in female sierra dome spiders, Linyphia litigiosa (Lynyphiidae). Animal Behaviour 41:343-360. Wolff, R.J. 1977. The cribellate genus Tengella (Araneae: Tengelli- dae?). Journal of Arachnology 5:139-144. Manuscript received 6 March 2007, revised 11 April 2008. 2008. The Journal of Arachnology 36:609-611 SHORT COMMUNICATION Snatching prey from the mandibles of ants, a feeding tactic adopted by East African jumping spiders Robert R. Jackson: International Centre for Insect Physiology and Ecology, PO Box 30772, Nairobi, Kenya & School of Biological Sciences, University of Canterbury, Private Bag 4800, Christchurch, New Zealand Kathryn Salm: School of Biological Sciences, University of Canterbury, Private Bag 4800 Christchurch, New Zealand Simon D, Pollard': Canterbury Museum, Rolleston Avenue, Christchurch 8013, New Zealand Abstract. Instances are documented of salticids robbing ants by adopting a specialized behavior pattern, “snatching.” The salticid positioned itself beside an ant column on the wall of a building, repeatedly fixating its gaze on different individual ants in the column and maintaining fixation on the ant by turning its body while the ant walked by. When close to an ant that was carrying prey, the salticid maneuvered about so that it was head on, grabbed hold of the prey using its chelicerae, and then rapidly pulled the prey out of the ant’s mandibles. Having secured the prey, the salticid moved away from the ant column to feed. All observations were made at Mbita Point, by the shore of Lake Victoria in western Kenya. The salticids were three species of Menemerus (Simon 1868): M. bivittatus (Dufour 1831 ), M. congoensis Lessert 1927 and an undescribed species, Menemerus sp. n. The ant species were from the genera Crematogaster (Lund 1831) and Camponotus (Mayr 1861 ). In all instances, the salticid was 2-6 mm in body length (juveniles of all three Menemerus species and adults of Menemerus sp. n). Prey items taken from ants were, in most instances, “lake flies” (adults of Chaoboridae and Chironomidae). Keywords: Salticidae, ants, predation, stealing prey, Chironomidae, Chaoboridae In the tropics, ants (Formicidae) are the dominant insects (Holldobler & Wilson 1990) and jumping spiders (Salticidae) are the dominant spiders (Coddington & Levi 1991), but we are only beginning to understand how salticids and ants interact (Nelson & Jackson 2005, 2006a, b; Nelson et al. 2006). Salticids are unique among spiders because of their complex eyes (Land 1969; Blest et al. 1990), exceptionally acute vision (Land & Nilsson 2002) and intricate vision-guided predatory behavior (Jackson & Pollard 1996; Harland & Jackson 2004). Most species in this large family (about 5,000 described species, Platnick 2008) appear to be active hunters that prey primarily on a variety of insects, but typically they do not prey on ants. It may not be surprising that many salticid species can detect ants by sight and then avoid coming close to them (Nelson & Jackson 2006c), particularly considering the formidable defences shown by ants (Blum 1981; Holldobler & Wilson 1990), including powerful mandibles, poison-injecting stings and formic-acid sprays, and the fact that ants are sometimes predators of salticids (Nelson et al. 2004). Yet there is a large minority of salticids (the “myrmecophagic species”) that selects ants as preferred prey (Li & Jackson 1996; Clark et al. 2000; Jackson & Li 2001; Huseynov et al. 2005) and one salticid species, Cosmophasis bitaeniata (Keyserling 1882), is known to combine chemical ant mimicry with myrmecophagy (Allan & Elgar 2001) (i.e., by mimicking the cuticular hydrocarbons of the Australian weaver ant, Oecophylla smaragdina (Fabricius 1775), C. bitaeniata gains entry to the weaver ant’s nest and feeds unmolested on the ant’s larvae). Here we revisit a different style of exploiting ants — robbing ants of objects they carry in their mandibles. This was First described by Bhattacharya (1936) who observed juveniles of Menemerus bivattatus (Dufour 1831) (formerly Marpissa melanognathus ) in India grabbing food out of the mandibles of Fire ants, Solenopsis geminata (Fabricius 1804). Our own observations show that this tactic, which we will call “snatching from ants” or just “snatching,” for short, is unique neither to India nor to M. bivittatus. The baseline information we provide here is a step toward later quantitative and experimental research concerned with this poorly understood foraging method. 1 Corresponding author. E-mail: simon. pollard@canterbury.ac.nz METHODS Menemerus is a large, well-deFined genus, probably with many of the African species yet to be described (Wesolowska 1999). Our observations were on M. bivittatus (Dufour 1831), M. congoensis (Lessert 1927) and a new, undescribed species, Menemerus sp. n. all three of which are common in East Africa (see Jackson 1986, 1999). Typical body lengths of adult females of each species are: M. bivittatus, 10 mm; M. congoensis, 7 mm; Menemerus sp. n. 5 mm. Voucher specimens of all species from this study (salticids, ants, and prey) have been deposited with the Florida State Collection of Arthropods in Gainesville and the National Museums of Kenya in Nairobi. Our study site was by the shore of Lake Victoria in western Kenya (Mbita Point, the Thomas Odhiambo Campus of the International Centre for Insect Physiology and Ecology). Mbita Point is 1200 m above sea level (0°25'S-0°30'S, 34 °10'E-35°15'E) and has a mean annual temperature of 21° C. In this habitat, midges (Diptera: Chironomidae & Chaoboridae), known locally as “lake flies,” are exceedingly abundant (Beadle 1981), often covering the walls of buildings. As midges have notoriously short life spans, lake-fly swarms quickly turn into enormous numbers of lake-fly corpses which are routinely scavenged by ants. We opportunistically observed Menemerus and other salticids when they were seen in the vicinity of ants on building walls. Whenever we saw a salticid persistently orienting toward an ant (identified to genus only), we continued observation for 30-60 min or until the salticid secured the prey. First we made about 30 preliminary observations of Menemerus sp. n. snatching lake flies from an unidentified species of Crematogaster (Lund 1831), but with no attempt made to identify the lake fly to family and no records kept concerning the salticid other than the species to which it belonged. We videotaped 10 of these preliminary observations for more detailed information about behavior. This was followed by observations (n = 98) that were more standardized with respect to the information we recorded. After each of these observations, we collected the salticid, the ant and the “prey” (i.e., object snatched from an ant). Salticids were identified to species, 609 610 THE JOURNAL OF ARACHNOLOGY ants to genus and lake flies to family. We also recorded whether the salticid was a juvenile or an adult, and we recorded whether adults were male or female. Earlier convention (Jackson and Li 2001) is adopted for indicating frequencies of occurrence: “usually,” “often,” “typically,” and “typical” indicate ca 80% or more. RESULTS Snatching from ants. —The three Menemerus species, as well as Evarcha culicivora (Wesolowska & Jackson 2003), Harmochirus brachiatus (Thorell 1877), Hasarius adansoni (Audouin 1826) and unidentified species of Hyllus (Koch 1846), Natta (Karsch 1879), Myrmarachne (MacLeay 1839), Plexippus (C.L. Koch 1846), and Thyene (Simon 1885), were common on the walls of buildings, with many other salticids present in smaller numbers. However, only the three Menemerus species were observed snatching from ants. In the records of salticids snatching front ants (n = 98), salticid body length varied from 2 mm ( n = 1, 1%, Menemerus sp. n.) to 6 mm (n = 1, 1%, M. bivittatus), with 6 (6%) being 3 mm, 27 (28%) being 4 mm and the majority (n = 63, 64%) being 5 mm. For the majority records of snatching from ants (n = 98), the salticid was Menemerus sp. n. (78, 80%), with 58 (74%) of these 78 records coming from adult females, 6 (8%) coming from adult males, and 14 (18%) coming from juveniles. For all records for Menemerus congoensis (n = 10, 10% of 98) and M. bivittatus ( n = 10, 10%), the salticid was a juvenile. The ants were undetermined species from the genus Crematogaster (n = 72, 73%) and Camponotus (Mayr 1861) (n = 26, 27%). Observed snatching sequences took place either in the morning (07:00-1 1:00 hours, n = 75) or in the late afternoon (17:00- 19:00 hours, n = 23). The objects snatched from ants were usually (90, 92% of 98) dead lake flies (body lengths: 5 to 10 mm) (chironomid, n = 77, 79% of 98) and chaoborid (n = 13, 13%). Besides lake flies, adult females of Menemerus sp. n. (body length 5 mm) were also observed snatching an ant egg (n = 1), a dead mayfly (Ephemeroptera, n = 1), a dead Crematogaster worker (n = 3) and what appeared to be plant material (/; = 3), with all of these objects being comparable to lake Hies in size. Menemerus sp. n subsequently ate the mayfly and the ant egg, but released and moved away from the plant material and the dead ant a few seconds after contact. There were also five instances in which Menemerus sp. n. (3 adult females and 2 juveniles) snatched a dead lake fly (not identified to family) from an ant and then, a few seconds later, released the lake fly and walked away. In all other instances, the salticid ate the lake fly it snatched from an ant. Behavioral sequences were similar irrespective of the different prey, ant genus, Menemerus species and, for Menemerus sp. n., whether the salticid was a juvenile, an adult male, or an adult female. Five behavioral stages were discerned: tracking, intercepting, attacking, retreating, and feeding. Tracking: a salticid positioned itself beside an ant column on the wall of a building, repeatedly fixating (i.e., aligning the gaze of the corneal lenses of its anterior-medial eyes) on different individual ants active in the column and maintaining fixation on each ant for 5 s or longer by continually turning its body while the ant walked by. The ant being tracked was usually carrying an object in its mandibles. The salticid usually remained 50-100 mm from the ant while tracking and stepped out of the way whenever an ant turned and moved in its direction. Intercepting: a spider that had been tracking suddenly began stepping about and maneuvering into position in front of the ant, effectively blocking the ant’s forward progress. This usually happened only a few seconds after the salticid was first seen tracking, as any longer delay usually resulted in the ant moving far away from the salticid. When intercepting, Menemerus usually took a veering path and approached the ant column 20-45° off from straight ahead of the targeted ant's forward trajectory. When Menemerus stepped in front of the ant, the ant either stopped momentarily before moving off in a different direction or it just slowed down and veered to the side, with Menemerus continuing to maneuver itself in front of the active ant. Attacking: during one of an ant’s momentary pauses when being intercepted or else while the ant was attempting to step out of the way, a spider suddenly extended its rear legs, moved its body 1-2 mm forward, brought its chelicerae into contact with an object in the ant’s mandibles and then immediately stepped a few millimeters backwards or to the side, pulling the object out of the ant’s mandibles. In all instances, the ant released the object when the salticid pulled away. Retreating: after extracting an object from the ant’s mandibles, the spider turned and rapidly walked away, usually not stopping until about 100-200 mm from the ant column. Feeding: the spider settled, usually in a space between bricks or in some other secluded location on the wall, and then proceeded to feed for 1-10 min. After feeding, the spider dropped the prey and walked away, after which it often returned to the ant column and stole another lake fly from the ants. As many as four lake flies were sometimes stolen in succession. There were about 10 instances each for M. congoensis and M. bivittatus , and more than 40 for Menemerus sp. n., in which we observed a salticid briefly tracking an ant that had empty mandibles, but we never saw a salticid intercept these ants. There were also about 30 instances in which Menemerus sp. n. briefly tracked, but then failed to intercept, as well as 9 instances of seeing a salticid track and then intercept an ant that was carrying an object other than a lake fly, but then move away without attacking (Menemerus sp. n., 5 ants carrying plant material and 2 carrying a dead conspecific ant worker; M. congoensis , 2 carrying dead conspecifics). DISCUSSION Bhattacharya (1936) provided minima! descriptive detail of snatching behavior. He did not indicate how many times he observed M. bivittatus snatching from ants and he referred to the objects M. bivittatus stole as simply “food and eggs” (ant, spider, and object sizes not indicated). Yet his observations on M. bivittatus in India appear to have been similar to ours: tracking, intercepting, attacking (pulling the prey out of the ant’s mandibles) and retreating with the prey before feeding. Bhattacharya (1936) also observed M. bivittatus adults, but not juveniles, stalking, capturing, and feeding on house Hies, Musca domestica (Linnaeus 1758) and he suggested that snatching prey from ants might be the primary foraging tactic of M. biviattatus juveniles. We hesitate to suggest that this is the primary tactic used by any of the active stages of any of the three Menemerus species we studied because we observed all stages of each of the three Menemerus species frequently capture and eat free (i.e., not in the mandibles of ants) living prey by practicing the stalk-and-leap routines that appear to be typical of many salticid species (Forster 1982; Richman & Jackson 1992; Jackson & Pollard 1996). Our observations suggest instead that snatching from ants is an alternative foraging tactic sometimes adopted by small individuals (i.e., individuals no more than 6 mm in body length) of Menemerus. For the smallest of the three Menemerus species we studied (i.e., Menemerus sp. n.), this included adults of both sexes as well as juveniles. However, for M. bivittatus and M. congoensis , this included only juveniles. Despite many salticid species being abundant at Mbita Point, the only salticids we saw snatching prey from ants were the three Menemerus species. These observations from East Africa, together with Bhattacharya's (1936) records from India, suggest that snatching from ants may be widespread among species of Menemerus living in ant-rich habitats. Further work is needed for determining whether this tactic is special to the genus Menemerus and for clarifying the selection pressures that might have favored the evolution of snatching behavior. JACKSON ET AL— SPIDERS SNATCH PREY FROM ANTS It is difficult to envisage a salticid needing an ant’s help overpowering inoffensive, soft-bodied lake flies and seeing building walls on the shore of Lake Victoria covered by lake flies does not suggest that finding lake flies is a pressing problem for which a salticid might need an ant’s assistance. We may be tempted by an image of salticids grazing on clumps of lake flies, rather like antelopes grazing on clumps of grass, but choosing and capturing a living lake fly may be far from effortless for a salticid. Time considerations may be important. Success for Menemerus during stalk-leap sequences may often depends on slowly moving close enough to gauge an accurate leap, with a targeted prey potentially nullifying the salticid’s efforts by flying away. Stalking sequences typically take several minutes, compared with the few seconds needed to intercept an ant. Decision making is another potential problem for a salticid. The image of unlimited lake-fly prey changes somewhat upon close examination. Many of the lake flies covering building walls are in fact already dead, but stray silk lines left by spiders hold dead lake flies in place in lifelike postures on the wall. A light breeze often makes the dead lake flies twitch and jiggle about. Menemerus and other salticids were often seen stalking these dead flies, leaping on them when close and then almost immediately releasing them, but there were only five instances in which we observed Menemerus sp. n immediately release and move away from a dead lake fly it had snatched from an ant. Perhaps one of the primary advantages of stealing from ants is that the salticid can rely on the ant to select lake (lies that are still fresh enough to be palatable. ACKNOWLDEGM ENTS We are especially grateful to Hans Herren, Christian Borgeimeister, Bart Knols, and Charles Mwenda at the International Centre for Insect Physiology and Ecology for the numerous ways in which they supported the research. Godfrey Sune, Stephen Alluoch, Silas Ouko Orima, and Jane Atieno provided invaluable technical assistance. We gratefully acknowledge support from the National Geographic Society (RRJ), the Royal Society of New Zealand (Marsden Fund; RRJ, SDP) and James Cook Fellowship (RRJ). For taxonomic assistance, we thank Wanda Wesolowska, G.B. Edwards, Roy Snelling and Arthur Harrison. LITERATURE CITED Allan, R.A. & M.A. Elgar. 2001. Exploitation of the green tree ant, Oecophylla smaragdina, by the salticid spider Cosmophasis bitaeniata. Australian Journal of Zoology 49:129-137. Beadle, L.C. 1981. The Inland Waters of Tropical Africa: An Introduction to Tropical Limnology. 2nd edition. Longman, New York. 457 pp. Bhattacharya, G.C. 1936. Observations of some peculiar habits of the spider ( Marpissa melanognathus). Journal of the Bombay Natural History Society 39:142-144. Blest, A.D., D.C. O’Carroll & M. Carter. 1990. Comparative ultrastructure of Layer 1 receptor mosaics in principal eyes of jumping spiders: the evolution of regular arrays of light guides. Cell Tissue Research 262:445-460. Blum, M.S. 1981. Chemical Defences of Arthropods. Academic Press, New York. 562 pp. Clark, R.J., R.R. Jackson & B. Cutler. 2000. Chemical cues from ants influence predatory behavior in Habrocestum pulex (Hentz), an ant-eating jumping spider (Araneae, Salticidae). Journal of Arachnology 28:299-341. Coddington, J.A. & H.W. Levi. 1991. Systematics and evolution of spiders (Araneae). Annual Review of Ecology and Systematics 22:565-592. 611 Forster, L.M. 1982. Vision and prey-catching strategies in jumping spiders. American Scientist 70:165-175. Harland, D.P. & R.R. Jackson. 2004. Portia perceptions: the Umwelt of an araneophagic jumping spider. Pp. 5-40. In Complex Worlds from Simpler Nervous Systems. (F.R. Prete, ed.). MIT Press, Cambridge, Massachusetts. Holldobler, B. & E.O. Wilson. 1990. The Ants. Belknap Press, Harvard University, Cambridge, Massachusetts. 746 pp. Huseynov, E.F., F.R. Cross & R.R. Jackson. 2005. Natural diet and prey-choice behaviour of Aelurillus muganicus (Araneae: Salt- icidae), a myrmecophagic jumping spider from Azerbaijan. Journal of Zoology, London 267:159-165. Jackson, R.R. 1986. Communal jumping spiders (Araneae, Salticidae) from Kenya. New Zealand Journal of Zoology 13:13-26. Jackson, R.R. 1999. Spider cities of Africa. New Zealand Science Monthly 10:10-11. Jackson, R.R. & D. Li. 2001. Prey-capture techniques and prey preferences of Zenodorus durvillei, Z. metallescens and Z. orbiculata, tropical ant-eating jumping spiders (Araneae: Salt- icidae) from Australia. New Zealand Journal of Zoology 28:299-341. Jackson, R.R. & S.D. Pollard. 1996. Predatory behavior of jumping spiders. Annual Review of Entomology 41:287-308. Land, M.F. 1969. Structure of the retinae of the principal eyes of jumping spiders (Salticidae: Dendryphantinae) in relation to visual optics. Journal of Experimental Biology 51:443-470. Land, M.F. & D.E. Nilsson. 2002. Animal Eyes. Oxford University Press, Oxford, UK. 221 pp. Li, D. & R.R. Jackson. 1996. Prey-specific capture behaviour and prey preferences of myrmecophagic and araneophagic jumping spiders (Araneae: Salticidae). Revue Suisse Zoologie, hors serie, 423-436. Nelson, X.J. & R.R. Jackson. 2005. Living with the enemy, jumping spiders that mimic weaver ants. Journal of Arachnology 33:813-819. Nelson, X.J. & R.R. Jackson. 2006a. Compound mimicry and trading predators by the males of sexually dimorphic Batesian mimics. Proceedings of the Royal Society of London (B) 273:367-372. Nelson, X.J., D. Li & R.R. Jackson. 2006b. Out of the frying pan and into the fire: a novel trade-off for Batesian mimics. Ethology 1 12:270-277. Nelson, X.J. & R.R. Jackson. 2006c. Vision-based innate aversion to ants and ant mimics. Behavioral Ecology 17:676-681. Nelson, X.J., R.R. Jackson, G.B. Edwards & A.T. Barrion. 2004. Predation by ants on jumping spiders (Araneae: Salticidae) in the Philippines. New Zealand Journal of Zoology 31:45-56. Nelson, X.J., R.R. Jackson & D. Li. 2006. Conditional use of honest signalling by a Batesian mimic. Behavioral Ecology 17:575-580. Platnick, N.I. 2008. The World Spider Catalogue. Version 7.0. American Museum of Natural History, New York. Online at http://research.amnh.-org/entomology/spiders/catalog81-87/index. html. Richman, D. & R.R. Jackson. 1992. A review of the ethology of jumping spiders (Araneae, Salticidae). Bulletin of the British Arachnological Society 9:33-37. Wesolowska, W. 1999. A revision of the spider genus Menemerus in Africa (Araneae: Salticidae). Genus 10:251-353. Manuscript received 8 August 2007, revised 21 April 2008. 2008. The Journal of Arachnology 36:612-614 SHORT COMMUNICATION Excretion behavior of adult female crab spiders Misumena vatia (Araneae, Thomisidae) Douglass H. Morse: Department of Ecology & Evolutionary Biology, Box G-W, Brown Elniversity, Providence, Rhode Island 02912, USA. E-mail: d_morse@brown.edu Abstract. Excreta potentially provide parasites or predators with information about the presence of hosts or prey; hence, vulnerable individuals experience strong selection to minimize danger from this source. Alternatively or additionally, excreta could alert potential prey to a spider’s presence. Adult female crab spiders Misumena vatia (Clerck 1757) exhibited a strong reluctance to excrete when retained under tightly confined conditions. Only 5% of regularly fed individuals (1% of total observations) excreted over observation periods of as many as 50 days while confined in 7-dram vials (5 cm high, 3 cm diameter). Individuals retained large amounts of excreta during this time. However, when released upon vegetation over two-thirds of them excreted within 5 min, after moving to the distal end of a leaf or petal such that the excreta fell below them onto lower vegetation or the substrate. In the field they showed little tendency to excrete close to their hunting sites. The ability to retain excreta should serve this relatively sedentary species well in situations where it suffers high rates of attack or may reveal its presence to potential prey. Keywords: Defecation, parasite avoidance, predator avoidance, retention behavior Many animals may experience risks in voiding excretory material, which likely enhance the possibility of alerting predators and parasites to their presence and increase the danger of disease. Alternatively or additionally, the presence of excreta could forewarn potential prey to a predator’s presence. Considering the importance of these factors, surprisingly little attention has focused on studies examining the responses of predatory invertebrates to host or prey waste products (Weiss 2006). In particular, workers have written little about excretion in spiders (Curtis & Carrel 2000) or other arachnids (Sato et al. 2003; Sato & Saito 2006). In fact, Curtis & Carrel (2000) believed their paper on excretion behavior by garden spiders Argiope aurantia Lucas 1833 (Araneidae) (referred to as defecation behavior by the authors) to be the first explicit study of its sort on a spider, although Tietjen (1980) reported on the nonrandom distribution of excreta under laboratory conditions in Mallos gregalis (Simon 1909) (Dictynidae). Throughout this paper I use the term “excretion” (excreta, excrete, etc.) to identify the materials passing through a spider’s anus, since the majority of this material consists of the products of post-assimilatory metabolic processes from the Malpi- ghian tubules, rather than undigested matter. In response to the risk of this material providing cues to their presence, potential prey or hosts might develop behavioral responses that minimize this threat, such as excreting away from their normal activities or retaining excreta indefinitely until they can safely void them. Taxa that show high fidelity to a site should experience particularly strong pressure to develop such tactics, as demonstrated by Weiss (2003, 2006) for caterpillars. Crab spiders Misumena vatia (Clerck 1757) (Thomisidae) typically occupy flowers as sit-and-wait predators of insects and in the process may often remain at hunting sites for several days at a time. Excretion in these areas could attract a large number of predators and parasites through either visual or olfactory cues provided by this material. Although various vertebrates are usually considered the most common predators of spiders, they only infrequently prey on Misumena in coastal Maine, where I conducted this study (Morse 1985, 2007). More important are other invertebrates, especially spiders, predatory wasps, and parasitoid wasps and flies (Morse 1988a, 1988b), some of which likely respond to olfactory cues. Here I characterize the excretion behavior of adult female Misumena retained for extended periods under confined conditions and then provided with sites that allowed them to dispose of their excreta some distance away from their hunting sites. I quantified the spiders’ frequency and size of excretion when confined to small vials and immediately following their release onto vegetation, both flowers in the laboratory and leafy vegetation in the field. I also report observations on the excretion patterns of free-ranging adult females in the field. In combination, these results allow me to test whether these spiders discriminate among potential excretion sites, whether the spiders’ ability to separate themselves from their excreta affects which sites they use for this purpose, and whether the site affects the size of the excretion. METHODS Adult female Misumena are medium-sized spiders that molt into the adult stage at 35-60 mg and may exceed 400 mg when they lay their single egg mass. These spiders have two large, raptorial anterior pairs of legs and two much smaller posterior pairs. They can change their color between white and yellow, and most have a prominent pair of deep red dorsolateral abdominal stripes. Misumena frequent old fields and roadsides in my study area (South Bristol, Lincoln County, Maine, USA), where they hunt on flowers for large prey. I collected 72 adult females from these sites in June and July for studies unrelated to this one. The design of that work dictated in part the types of observations that I could make for this study. I kept the spiders in 7-dram vials (5 cm tall, 3 cm diameter) at ambient temperature and light regimes and fed them a moth (Noctuidae, Geometridae) or large fly (Syrphidae, Muscidae) every other day. Adult female spiders grew rapidly on this diet and did not require supplementary liquids. In the process I retained individuals for several days to well over one month. Retention times of the spiders varied in accordance with their mass upon capture and how rapidly they gained mass up to the point of egg laying. I recorded excretions when feeding the spiders and cleaning their vials after feeding or excretion. I did not start recording retention times of excreta by the spiders until they had been in the vials for two days to ensure that all individuals were in a similar hunger state. These spiders will usually take a large prey item every other day (Morse & Fritz 1982). Numbers of drops of excreta were counted whenever possible. For the laboratory observations, I released female Misumena from their vials onto a flower, either an oxeye daisy Chrysanthemum leucanthemum or black-eyed Susan Rudebeckia hirta, and observed 612 MORSE— CRAB SPIDER EXCRETION 613 their behavior. I then counted all excretions produced within the next five minutes. Previous observations had demonstrated that the females often excreted immediately after release on a fiower, particularly if 1 had retained them for several days before release (D.H. Morse, unpub. obs.). I placed another group of previously confined adult females in the field on young, non-flowering milkweed Asclepias syriaca plants, sites previously recognized as favored nesting places (Morse 1985). Upon releasing spiders onto the plants, I observed these individuals for five minutes to determine whether they would excrete during that period, since earlier unrelated observations had established that they often excreted within this time. I also tested the frequency with which free-ranging adult female Misumena excreted in conspicuous hunting sites on wild marjoram Origanum vulgare over periods as long as 17 days. These marjoram stems averaged 0.5 m in height and grew densely at the test site, which contained several hundred flowering stems. They bore terminal rounded panicles composed of multiple small pinkish-purple flowers. Immediately below their inflorescences marjoram stems bear dense ovate leaves, such that the hunting sites in marjoram occur in the top of a dense canopy of flowers and leaves. This situation made it difficult for spiders to excrete from their hunting sites without soiling nearby vegetation. RESULTS Characteristics of behavior and excreta. -Misumena exhibited distinctive excretion behavior: individuals moved to the tip of a petal or leaf, raised themselves on their two pairs of large forelimbs, the two smaller pairs of posterior limbs usually not contacting the substrate at this time, and then released varying numbers of drops of a whitish liquid with dark brown flecks that quickly dried in the air to a dirty light brown. Upon release onto the flowers the spiders typically commenced excretion behavior quickly, often within the first 30 seconds. In the laboratory, excretions made from the tips of flower petals fell onto the substrate below; in the field they most often completely cleared the plant in question, landing on the grass in the substrate. When voided from milkweed leaves in the field, the excreta most often landed on another leaf of the plant below the excretion site. If permitted to climb to the rim of their vial the spiders readily excreted from there as well, exhibiting the same behavior as seen on the petals and leaves (D.H. Morse, pers. observ.). During large excretions in which the spiders voided many drops, they released most of these drops in a nearly constant stream, so that my counts of these drops were approximate. These excretions averaged 5.0 ± 0.97% (± SE, n = 6) of the previous body mass (D.H. Morse, unpub. data). The spiders distinctly spaced these drops in smaller excretions. Tendency to excrete. — I recorded only four excretions in the vials during the feeding and cleaning sessions that took place every second or third day. These involved 72 spiders and 335 observations of spiders at these sessions, with spiders present for 115 such sessions (5.6% of the spiders and 1.2% of the total observations showed excretion). I did not retain most individuals long enough to obtain probable maximum retention times of excreta, but individuals regularly refrained from excreting for up to one month or more, with a maximum of 47 days. Unfortunately I failed to record which individuals excreted, but even if one assumes that the four longest- remaining individuals (47, 39, 39, 35 days) excreted, thereby accounting for the four excretions recorded during this period, seven individuals retained their excreta for over 30 days [34 (2), 33 (4), 32 (1)]. Thus, individuals could routinely retain their excreta for long periods. In laboratory observations, 29 (67.4%) individuals excreted during the five-minute period after release from the vial, and 14 failed to excrete at this time. This result differed highly significantly from the number expected from the spiders’ behavior in the vials, which would predict zero or one excretion (G = 40.51, df = I , P < 0.001 in a G-test for goodness of fit). None of these individuals excreted in subsequent minutes. When released on the milkweed plants, 28 (73.7%) individuals excreted within five minutes, and 10 did not excrete, a highly significant difference, using the same rationale as the previous test ( G = 55.39, df = 1, P < 0.001, same test). Size of excretions. Excretions, measured as drops of liquid, differed widely in volume, probably a consequence of how long individuals had retained this material. In the laboratory sessions excretions averaged (± SE) 6.6 ± 1.50 drops, range = 1 to 26 (n = 22 observations); on the milkweeds they averaged 9.0 ± 1.62 drops, range = 1 to 24 (n = 18 observations). Excretions at release in the field significantly exceeded those during laboratory sessions ( U = 130.5, P < 0.03 in a one-tailed Mann-Whitney U test). On average, retention of excreta at release should exceed those recorded in the laboratory. The size of excretions at release was positively related to the time that spiders had retained this material (R2 = 0.41, n = 17. P < 0.01). In contrast, no relationship occurred between the size of excretions of individuals in the laboratory sessions and their retention times (R2 = < 0.01, n = 15, P > 0.9, same test). Behavior in the field. — During six censuses run every third day on marjoram, I made 45 observations of the free-ranging spiders out of a possible 120 (number of spiders released X number of counts). I observed a maximum of 1 1 spiders during a census, although recording 18 of them during one census or another (mean ± SE = 7.5 ± 0.96 individuals). As a result, these spiders might have spent as much as 62.5% (75 of 120 possible observations) of their time away from hunting sites, and a minimum of 27.5%, based on unrecorded individuals later found in the flowers (33 instances). These absences would provide ample opportunity for the spiders to excrete unnoticed. In the process of this census 1 failed to find a single excretion in the vicinity of a hunting site. DISCUSSION Adult female Misumena often retain their excreta for long periods under experimental conditions, a trait that should help to facilitate their relatively sedentary behavior. Since these spiders often remain for several days at a time on a superior hunting site (Morse & Fritz 1982), they should experience strong selection to void their excreta carefully. Seldom if ever did the location of such an individual become conspicuous in the field (at least to the human eye) as a result of their disposition of excreta (D.H. Morse, unpub. observ.), including the explicit observations reported here. Spiders possess a large stercoral pocket (cloacal chamber) with a muscular sphincter that allows them to store large amounts of excreta (Seitz 1987). The clear difference in relationship between size of excretion and time confined accords with the spiders excreting more regularly in the field than under confined laboratory conditions. Under these circumstances the individuals tested upon release in the field would have gone longer without excreting than those measured earlier in the laboratory, and hence, since fed regularly, would have accumulated significantly more excreta. Less likely, they might simply void less prior to the experiments, though I have no basis to support this alternative. Curtis & Carrel (2000) reported that garden spiders fed mealworms excreted over twice a day under otherwise normal field conditions. The spiders hunting on marjoram would have had little opportu- nity to excrete without soiling nearby vegetation if they had remained on their hunting sites. They might excrete low in the vegetation without my detecting them. Such behavior would match other observations suggesting that the spiders excrete more readily at sites where the excreta fall far below their hunting areas than where the excreta would fall in their midst (D.H. Morse, unpub. observ.). Dropping excreta from their immediate vicinity to the vegetation beneath them should make the spiders more conspicuous to other 614 THE JOURNAL OF ARACHNOLOGY animals on or near the substrate than to those in the canopy or above it. However, dropping their excreta away from their canopy-level hunting sites should make the spiders less vulnerable to most winged attackers, probably the most important threats to spiders in the leafy canopy. When they venture down onto the grassy substrate they expose themselves to attacks from such predators as meadow voles Microtus pennsylvanicus and garter snakes Thamnophis sirtalis (Morse 1985). In the canopy the egg predator Trychosis cyperia (Ichneumo- nidae) is their most important threat (Morse 1988b). In all of my observations on Misumena , I have never seen them preyed upon by birds (Morse 2007); however, the spider wasp Dipogon sayi (Pompilidae), rare in the study areas, takes a very occasional small adult female (two observations: D.H. Morse, unpub. data), and large sphecid wasps (Sphecidae), also uncommon in the study areas, are other potential predators, especially of penultimates (Morse 2007). 1 have little information on the role of disease or internal parasites, factors that should also favor careful disposition of excreta. I have twice reared horsehair worms (Gordioida) from adult females (D.H. Morse, unpub. data), but these events are relatively rare, since the two records come from a sample of several thousand females collected in the field as adults or penultimates. Tietjen (1980) reported little sign of bacterial or fungal growth about excreta of Mallos gregalis. Behavioral traits, however, may play an important role in controlling levels of parasitism of other animals (e. g., Hart 1992; Ezenwa 2004). Other sparse information on the excretion behavior of spiders suggests that they minimize the apparency of their excreta at their norma! hunting level in the vegetation. Curtis and Carrel (2000) noted that the garden spider often leaves its web to excrete, and that it generally does so at night; however, it excretes under its web, consistent with their impression that its major predators are birds and predatory wasps. Mallos gregalis concentrated its excreta in parts of an experimental enclosure that it used least frequently (Tietjen 1980). Bonnet (1930) reported that the fishing spider Dolomedes fimbriatus (Pisauridae) forcibly cast its excreta out from as far as 3-4 cm from its body. It also frequently excreted before jumping into the water, which could divert a would-be predator, as suggested by Seitz (1987). Although typically presented in the context of predator avoidance, some results may equally well minimize apparency of the spiders to prospective prey, as described by Brown et al. (1995) for pike-minnow interactions. I am unaware of any instances in which excretion patterns of spiders can be unequivocally attributed to minimizing apparency to prey, though this relationship might obtain in many instances, perhaps simultaneously with predator or parasite avoid- ance. Taken together, these observations, in combination with those of Misumena reported here, all suggest that distinctive patterns of excretion behavior may be widespread, but frequently ignored, among spiders. ACKNOWLEDGMENTS I thank K. J. Eckelbarger, T. E. Miller, L. Healy, and other staff members of the Darling Marine Center of the University of Maine for facilitating work on the premises; M. Weiss and an anonymous reviewer for comments on the manuscript; and J. Rovner for enlightening commentary on the appropriate use of the words “excretion” and “defecation” in spider biology. Voucher specimens from this study were deposited in the American Museum of Natural History, New York. LITERATURE CITED Bonnet, P. 1930. La mue, 1’autotomie et la regeneration chez les araignees, avec une etude des Dolomedes d’Europe. Bulletin de la Societe d’histoire naturelle de Toulouse 59:237-700. Brown, G.E., D.O. Chivers & RJ.F. Smith. 1995. Localized defecation by pike: a response to labeling by cyprinid alarm pheromone. Behavioral Ecology and Sociobiology 36:105-110. Curtis, J.T. & J.E. Carrel. 2000. Defaecation behaviour of Argiope aurantia (Araneae: Araneidae). Bulletin of the British Arachno- logical Society 11:339-342. Ezenwa, V.O. 2004. Selective defecation and selective foraging: antiparasite behavior in wild ungulates? Ethology 110:851-862. Hart, B.L. 1992. Behavioral adaptations to parasitism: an ethological approach. Journal of Parasitology 78:256-265. Morse, D.H. 1985. Nests and nest-site selection of the crab spider Misumena vatia (Araneae, Thomisidae) on milkweed. Journal of Arachnology 13:383-390. Morse, D.H. 1988a. Relationship between crab spider Misumena vatia nesting success and earlier patch-choice decisions. Ecology 69:1970-1973. Morse, D.H. 1988b. Interactions between the crab spider Misumena vatia (Clerck) (Araneae) and its ichneumonid egg predator Trychosis cyperia Townes (Hymenoptera). Journal of Arachnology 16:132-135. Morse, D.H. 2007. Predator upon a Flower. Harvard University Press, Cambridge, Massachusetts. 392 pp. Morse, D.H. & R.S. Fritz. 1982. Experimental and observational studies of patch-choice at different scales by the crab spider Misumena vatia. Ecology 63:172-182. Sato, Y. & Y. Saito. 2006. Nest sanitation in social spider mites: interspecific differences in defecation behavior. Ethology 112:664-669. Sato, Y„ Y. Saito & T. Sakagami. 2003. Rules for nest sanitation in a social spider mite, Schizotetranychus miscanthi Saito (Acari: Tetranychidae). Ethology 109:713-724. Seitz. K.-A. 1987. Excretory organs. Pp. 239-248. In Ecophysiology of Spiders. (W. Nentwig, ed.). Springer- Verlag, Berlin. Tietjen, W.J. 1980. Sanitary behavior by the social spider Mallos gregalis (Dictynidae): distribution of excreta as related to web density and animal movements. Psyche 87:59-73. Weiss, M.R. 2003. Good housekeeping: why do shelter-dwelling caterpillars fling their frass? Ecology Letters 6:361-370. Weiss, M.R. 2006. Defecation behavior and ecology of insects. Annual Review of Entomology 51:635-661. Manuscript received 7 December 2007, revised 26 May 2008. 2008. The Journal of Arachnology 36:615-616 INSTRUCTIONS TO AUTHORS SUBMITTING ARTICLES TO THE JOURNAL OF ARACHNOLOGY (instructions last revised September 2008) General: Manuscripts must be in English and should be prepared in general accordance with the current edition of the Council of Biological Editors Style Manual unless instructed otherwise below. Use the active voice throughout. Authors should consult a recent issue of the Journal of Arachnology for additional points of style. Manuscripts longer than three printed journal pages (12 or more double-spaced manuscript pages) should be prepared as Feature Articles, shorter papers as Short Communications. One invited Review Article will be published in the first issue of each year at the discretion of the editors. Suggestions for review articles may be sent to the Managing Editor. Unsolicited review articles are also wel- comed. All review articles will be subject to the same review process as other submissions. Submission: Submissions must be sent electronically in Microsoft Word (or Word compatible) format. Large plates and figures may be sent as pdf or jpg files. Send submissions to the Managing Editor of the Journal of Arachnology. Douglass H. Morse, Managing Editor, Hermon Carey Bumpus Professor of Biology Emeritus, Department of Ecology & Evolutionary Biology, Box G-W, Brown University, Providence, RI 02912 USA [Telephone: 401-863-3152; Fax: 401-863-2166; E-mail: d_morse@brown.edu] The Managing Editor will acknowledge receipt of the manuscript, assign it a manuscript number and forward it to an Associate Editor for the review process. Correspondence relating to manuscripts should be directed to the Associate Editor and should include the manuscript number. If the manuscript is accepted, the author will be asked to submit the final copy electronically to the Associate Editor. Submission of final illustrations is detailed below. Authors are expected to return revisions promptly. Revised manuscripts that are not returned in a reasonable time period (no longer than six months for minor revisions and one year for major revisions) will be considered new submissions. Voucher Specimens: Specimens of species used in your research should be deposited in a recognized scientific institution. All type material must be deposited in a recognized collection/institution. FEATURE ARTICLES Title page. — The title page includes the complete name, address, and telephone number of the corresponding author; a FAX number and electronic mail address if available; the title in upper and lower case, with no more than 65 characters and spaces per line in the title; each author’s name and address; and the running head. Running head. — The author’s surname(s) and an abbreviat- ed title should be typed in all capital letters and must not exceed 60 characters and spaces. The running head should be placed near the top of the title page. Abstract. — The heading should be placed at the begin- ning of the first paragraph set off by a period. A second abstract, in a language appropriate to the nationality of the author(s) or geographic region(s) emphasized, may be included. Keywords. — Give 3-5 appropriate keywords or phrases following the abstract. Keywords should not duplicate words in the title. Text.- Double-space text, tables, legends, etc. throughout. Three levels of heads are used. • The first level (METHODS, RESULTS, etc.) is typed in capitals and centered on a separate line. • The second level head begins a paragraph with an in- dent and is separated from the text by a period and a dash. • The third level may or may not begin a paragraph but is italicized and separated from the text by a colon. Use only the metric system unless quoting text or referencing collection data. If English measurements are used when referencing collection data, then metric equivalents should also be included parenthetically. All decimal fractions are indicated by a period (e.g., -0.123). Include geographic coordinates for collecting locales if possible. Citation of references in the text: Cite only papers already published or in press. Include within parentheses the surname of the author followed by the date of publication. A comma separates multiple citations by the same author(s) and a semicolon separates citations by different authors, e.g., (Smith 1970), (Jones 1988; Smith 1993), (Smith & Jones 1986, 1987; Jones et al. 1989). Include a letter of permission from any person who is cited as providing unpublished data in the form of a personal communication. Citation of taxa in the text: Include the complete taxonomic citation for each arachnid taxon when it first appears in the manuscript. For Araneae, this information can be found online at http://research.amnh.org/entomology/spiders/cata- log/INTR02.html. For example, Araneus diadematus Clerck 1757. Citations for scorpions can be found in the Catalog of the Scorpions of the World (1758-1998) by V. Fet, W.D. Sissom, G. Lowe & M.E. Braunwalder. Citations for pseudoscorpions can be found in the Catalogue of the Pseudoscorpionida by M.S. Harvey. Citations for some species of Opiliones can be found in the Annotated Catalogue of the Laniatores of the New World (Arachnida, Opiliones) by A.B. Kury. Citations for other arachnid orders can be found in Catalogue of the Smaller Arachnid Orders of the World by Mark S. Harvey. Literature Cited section. — Use the following style and formatting exactly as illustrated; include the full unabbre- viated journal title. Personal web pages should not be included in Literature Cited. These can be cited within the text as (John Doe, pers. website) without the URL. Institutional websites may be included in Literature Cited. 615 616 THE JOURNAL OF ARACHNOLOGY Carico, J.E. 1993. Trechaleidae: a “new” American spider family. Pp. 305. In Proceedings of the Ninth International Congress of Arachnology, Panama 1983. (W.G. Eberhard, Y.D. Lubin & B.C. Robinson, eds.). Smithsonian Institu- tion Press, Washington, D.C. Huber, B.A. & W.G. Eberhard. 1997. Courtship, copulation, and genital mechanics in Physocyclus globosus (Araneae, Pholcidae). Canadian Journal of Zoology 74:905-918. Krafft, B. 1982. The significance and complexity of commu- nication in spiders. Pp. 15-66. In Spider Communications: Mechanisms and Ecological Significance. (P.N. Witt & .I.S. Rovner, eds.). Princeton University Press, Princeton, New Jersey. Platnick, N.I. 2006. The World Spider Catalog, Version 7.0. American Museum of Natural History, New York. On- line at http://research.amnh.org/entomology/spiders/catalog/ INTR01.html Roewer, C.F. 1954. Katalog der Araneae, Volume 2a. Institut Royal des Sciences Naturelles de Belgique, Bruxelles. 923 pp. Footnotes. — Footnotes are permitted only on the first printed page to indicate current address or other information concerning the author. All footnotes are placed together on a separate manuscript page. Tables and figures may not have footnotes. Taxonomic articles. — Consult a recent taxonomic article in the Journal of Arachnology for style or contact the Subject Editor for Taxonomy and Systematics. Papers containing original descriptions of focal arachnid taxa should be listed in the Literature Cited section. Tables. — Each table, with the legend above, should be placed on a separate manuscript page. Only horizontal lines (no more than three) should be included. Tables may not have footnotes; instead, include all information in the legend. Illustrations. Original illustrations should be sent electron- ically when the manuscript is submitted, preferably in tiff or jpeg format. Distribution maps should be considered figures and numbered consecutively with other figures. (Authors wishing to submit figures as hard copies should contact the Editor-in-Chief for specifications.) At the submission and review stages, the resolution standards may be low as long as editors and reviewers can view figures effectively. Final illustrations must be submitted to the Editor-in-Chief, typically by e-mail or on a CD, to ensure that the electronic versions meet publication standards and that they match the printed copy. All figures should be at least 4 inches wide, but no larger than a sheet of letter-size paper with 1-inch margins all around. The resolution should be at least 300 dpi (or ppi) for halftone or color figures and 1200 dpi for line drawings. A guide to the Digital Art Specs for Allen Press is posted online at: http://www2.allenpress.com/allen_press/apguides/Digital_ Art_Spec.pdf. To determine if your electronic figures adhere to the Allen Press specifications, you can also go to http:// verifig.allenpress.com, type in the password, allenpresscmy, and follow the instructions. Color plates can be printed, but the author must assume the full cost paid in advance, currently about $1100 US per color plate. Alternatively, an author can opt to have a figure printed in black and white, but for $30/ page have that figure appear in color in the online version of the journal. Most figures will be reduced to single-column width (9 cm, 3.5 inches), but large plates can be printed up to two-columns width (18 cm, 7 inches). Address all questions concerning illustrations to the Editor- in-Chief of the Journal of Arachnology. James E. Carrel, Editor-In-Chief, Division of Biological Sciences, 209 Tucker Hall, University of Missouri-Columbia, Columbia, MO 65211- 7400, USA [Telephone: 573-882-3037; FAX: 573-882-0123; E- mail: carrelj@missouri.edu] Legends for illustrations should be placed together on the same page(s) and separate from the illustrations. Each plate must have only one legend, as indicated below: Figures 1-4. A-us x-us , male from Timbuktu. 1, Left leg; 2, Right chelicera; 3, Dorsal aspect of genitalia; 4, Ventral aspect of abdomen. Scale = 1.0 mm. The following alternate Figure numbering is also accept- able: Figures la-e. A-us x-us , male from Timbuktu, a. Left leg; b. Right chelicerae; c. Dorsal aspect of genitalia; d. Ventral aspect of abdomen. Scale = 1.0 mm. Assemble manuscript. The manuscript should appear in separate sections or pages in the following sequence; title page, abstract, text, footnotes, tables with legends, figure legends, figures. If possible, send entire manuscript, including figures, as one Microsoft Word document. If figures or plates are large, please separate them from the text and send them as a pdf or jpg file. Page charges, proofs and reprints. — Page charges are voluntary, but non-members of AAS are strongly encouraged to pay in full or in part for their article ($75 / journal page). The author will be charged for changes made in the proof pages. Hard copy or pdf reprints are available only from Allen Press and should be ordered when the author receives the proof pages. Allen Press will not accept reprint orders after the paper is published. The Journal of Arachnology also is available through www.bioone.org and www.jstor.org. There- fore, you can download the PDF version of your article from one of these sites if you or your institution is a member. PDFs of articles older than one year will be made freely available from the AAS website. SHORT COMMUNICATIONS Short Communications are usually limited to three journal pages, including tables and figures (11 or fewer double-spaced manuscript pages including Literature Cited; no more than 2 small figures or tables). Internal headings (METHODS, RESULTS, etc.) are omitted. Short communications must include an abstract and keywords. COVER ARTWORK Authors are encouraged to send quality photographs (preferably in color) to the editor-in-chief to be considered for use on the cover. Images should be at least 300 dpi. CONTENTS Journal of Arachnology Volume 36 Featured Articles Number 3 A redescription of Varacosa apothetica (Wallace) (Araneae, Lycosidae) by Jamin M. Dreyer 487 Description of Zabius gaucho (Scorpiones, Buthidae), a new species from southern Brazil, with an update of the generic diagnosis by Luis E. Acosta, Denise M. Candido, Erica H. Backup & Antonio D. Brescovit 491 Revision of the spider genus Taira (Araneae, Amaurobiidae, Amaurobiinae) by Zhi-Sheng Zhang, Ming-Sheng Zhu & Da-Xiang Song 502 Subtle pedipalp dimorphism: a reliable method for sexing juvenile spiders by Nagissa Mahmoudi, Maria Modanu, Yoni Brandt & Maydianne C.B. Andrade 513 Activity pattern of the Neotropical harvestman Neosadocus maximus (Opiliones, Gonyleptidae): sexual and temporal variations by Francini Osses,Tais M. Nazareth & Glauco Machado 518 First male sperm precedence in multiply-mated females of the cooperative spider Anelosimus studiosus (Araneae, Theridiidae) by Thomas C. Jones & Patricia G. Parker 527 Frequency and consequences of damage to male copulatory organs in a widow spider by Michal Segoli, Yael Lubin & Ally R. Harari 533 Molting interferes with web decorating behavior in Argiope keyserlingi (Araneae, Araneidae) by Andre Walter, Mark A. Elgar, P. Bliss & Robin F. A. Moritz 538 Ontogenetic changes in web architecture and growth rate of Tengella radiata (Araneae, Tengellidae) by Gilbert Barrantes & Ruth Madrigal-Brenes 545 Does the microarchitecture of Mexican dry forest foliage influence spider distribution? by Pablo Corcuera, Maria Luisa Jimenez & Pedro Luis Valverde 552 Microhabitat preferences for the errant scorpion, Centruroides vittatus (Scorpiones, Buthidae) by C. Neal McReynolds 557 Observations on phenology and overwintering of spiders associated with apple and pear orchards in south-central Washington by Eugene R. Miliczky, David R. Horton & Carrol O. Calkins 565 Distribution of Geraeocormobius sylvarum (Opiliones, Gonyleptidae): range modeling based on bioclimatic variables by Luis E. Acosta 574 Ecology and web allometry of Clitaetra irenae, an arboricolous African orb-weaving spider (Araneae, Araneoidea, Nephilidae) by Matjaz Kuntner, Charles R. Haddad, Gregor Aljancic & Andrej Blejec 583 Differential survival of Geolycosa xera archboldi and G. hubbelli (Araneae, Lycosidae) after fire in Florida scrub by James E. Carrel 595 Book Review Harvestmen: the Biology of Opiliones. Edited by Ricardo Pinto-da-Rocha, Glauco Machado and Gonzalo Giribet. 2007. Harvard University Press, Cambridge, Massachusetts. 597 pp. ISBN 13:978-0-674-02343-7. US$125. by William A. Shear 600 Short Communications A new species of Xysticus (Araneae, Thomisidae) from Alberta, Canada by Charles D. Dondale 601 An easy method for handling the genus Phoneutria (Araneae, Ctenidae) for venom extraction by Leandro F. Garcia, Luiz Henrique A. Pedrosa & Denise R. B. Rosada 604 Courtship behavior and copulation in Tengella radiata (Araneae, Tengellidae) by Gilbert Barrantes 606 Snatching prey from the mandibles of ants, a feeding tactic adopted by East African jumping spiders by Robert R. Jackson, Kathryn Salm & Simon D. Pollard 609 Excretion behavior of adult female crab spiders Misumena vatia (Araneae, Thomisidae) by Douglass H. Morse. ... 612 INSTRUCTIONS TO AUTHORS 615 USERNAME: emerton08 PASSWORD: therid08 SMITHSONIAN INSTITUTION LIBRARIES 3 9088 01445 0076