S H ( l ,-A3- V5S U.S. Department of Commerce Volume 108 Number 4 October 2010 Fishery Bulletin U.S. Department of Commerce Gary Locke Secretary of Commerce National Oceanic and Atmospheric Administration Jane Lubchenco, Ph.D. Administrator of NOAA National Marine Fisheries Service Eric C. Schwaab Assistant Administrator for Fisheries Scientific Editor Richard D. Brodeur, Ph.D. Associate Editor Julie Scheurer National Marine Fisheries Service Northwest Fisheries Science Center 2030 S. Marine Science Dr. Newport, Oregon 97365-5296 Managing Editor Sharyn Matriotti National Marine Fisheries Service Scientific Publications Office 7600 Sand Point Way NE Seattle, Washington 98115-0070 The Fishery Bulletin (ISSN 0090-0656) is published quarterly by the Scientific Publications Office, National Marine Fisheries Service, NOAA, 7600 Sand Point Way NE, BIN C15700, Seattle, WA 98115-0070. Periodicals postage is paid at Seattle, WA. POSTMASTER: Send address changes for subscriptions to Fish- ery Bulletin, Superintendent of Docu- ments, Attn.: Chief, Mail List Branch, Mail Stop SSOM, Washington, DC 20402- 9373. Although the contents of this publica- tion have not been copyrighted and may be reprinted entirely, reference to source is appreciated. The Secretary of Commerce has deter- mined that the publication of this peri- odical is necessary according to law for the transaction of public business of this Department. Use of funds for printing of this periodical has been approved by the Director of the Office of Management and Budget. For sale by the Superintendent of Documents, U.S. Government Printing Office, Washington, DC 20402. Subscrip- tion price per year: $36.00 domestic and $50.40 foreign. Cost per single issue: $21.00 domestic and $29.40 foreign. See back for order form. Editorial Committee John Carlson Kevin Craig Jeff Leis Rich McBride Rick Methot Adam Moles Frank Parrish Dave Somerton Ed Trippel Mary Yoklavich National Marine Fisheries Service, Panama City, Florida Florida State University, Tallahassee, Florida Australian Museum, Sydney, New South Wales, Australia National Marine Fisheries Service, Woods Hole, Massachusetts National Marine Fisheries Service, Seattle, Washington National Marine Fisheries Service, Auke Bay, Alaska National Marine Fisheries Service, Honolulu, Hawaii National Marine Fisheries Service, Seattle, Washington Department of Fisheries and Oceans, St. Andrews, New Brunswick, Canada National Marine Fisheries Service, Santa Cruz, California Fishery Bulletin web site: www.fisherybulletin.noaa.gov The Fishery Bulletin carries original research reports and technical notes on investigations in fishery science, engineering, and economics. It began as the Bulletin of the United States Fish Commission in 1881; it became the Bulletin of the Bureau of Fisheries in 1904 and the Fishery Bulletin of the Fish and Wildlife Service in 1941. Separates were issued as documents through volume 46; the last document was No. 1103. Beginning with volume 47 in 1931 and continuing through volume 62 in 1963, each separate appeared as a numbered bulletin. A new system began in 1963 with volume 63 in which papers are bound together in a single issue of the bulletin. Beginning with volume 70, number 1, January 1972, the Fishery Bulletin became a periodical, issued quarterly. In this form, it is available by subscription from the Superintendent of Documents, U.S. Government Printing Office, Washington, DC 20402. It is also available free in limited numbers to libraries, research institutions, State and Federal agencies, and in exchange for other scientific publications. U.S. Department of Commerce Seattle, Washington Volume 108 Number 4 October 2010 Fishery Bulletin Contents Articles 365-381 Jones, Ashlee A., Norman G. Hall, and lan C. Potter Species compositions of elasmobranchs caught by three different commercial fishing methods off southwestern Australia, and biological data for four abundant bycatch species 382-392 Hurst, Thomas P., Benjamin J. Laurel, and Lorenzo Ciannelli Ontogenetic patterns and temperature-dependent growth rates in early life stages of Pacific cod ( Gadus macrocephcilus) The National Marine Fisheries Service (NMFS) does not approve, recommend, or endorse any proprie- tary product or proprietary material mentioned in this publication. No reference shall be made to NMFS, or to this publication furnished by NMFS, in any advertising or sales promotion which would indicate or imply that NMFS approves, recom- mends, or endorses any proprietary product or proprietary material mentioned herein, or which has as its purpose an intent to cause directly or indirectly the advertised product to be used or purchased because of this NMFS publication. The NMFS Scientific Publications Office is not responsible for the con- tents of the articles or for the stan- dard of English used in them. 393-407 Bollens, Stephen M., Rian vanden Hooff, Mari Butler, Jeffery R. Cordell, and Bruce W. Frost Feeding ecology of juvenile Pacific salmon ( Oncorhynchus spp.) in a northeast Pacific fjord: diet, availability of zooplankton, selectivity for prey, and potential competition for prey resources 408-419 DeCelles, Gregory R., and Steven X. Cadrin Movement patterns of winter flounder ( Pseudopleuronectes americanus) in the southern Gulf of Maine: observations with the use of passive acoustic telemetry 420-432 Heupel, Michelle R., Ashley J. Williams, David J. Welch, Campbell R. Davies, Ann Penny, Jacob P. Kritzer, Ross J. Marriott, and Bruce D. Mapstone Demographic characteristics of exploited tropical lutjanids: a comparative analysis 433-441 Graves, John E., and Andrij Z. Horodysky Asymmetric conservation benefits of circle hooks in multispecies billfish recreational fisheries: a synthesis of hook performance and analysis of blue marlin ( Makaira nigricans) postrelease survival II Fishery Bulletin 108(4) 442-449 Laidig, Thomas E. Influence of ocean conditions on the timing of early life history events for blue rockfish ( Sebastes mystinus) off California 450-465 Dunton, Keith J., Adrian Jordaan, Kim A. McKown, David O. Conover, and Michael G. Frisk Abundance and distribution of Atlantic sturgeon (Acipenser oxyrinchus) within the Northwest Atlantic Ocean, determined from five fishery-independent surveys 466-477 Hart, Ted D., Julia E. R. Clemons, W. Waldo Wakefield, and Selina S. Heppell Day and night abundance, distribution, and activity patterns of demersal fishes on Heceta Bank, Oregon 478-490 Simms, Jeffrey R., Jay R. Rooker, Scott A. Holt, G. Joan Holt, and Jessica Bangma Distribution, growth, and mortality of sailfish Ustiophorus platypterus) larvae in the northern Gulf of Mexico 491 Acknowledgment of reviewers 492 List of titles 494 List of authors 495 Subject index 498 Guidelines for authors Subscription form (inside back cover) 365 Species compositions of elasmobranchs caught by three different commercial fishing methods off southwestern Australia, and biological data for four abundant bycatch species Ashlee A. Jones (contact author)1 Norman G. HalS1 Ian C. Potter1 Email address for contact author: ashlee.|ones@murdoch. edu.au 1 Centre for Fish and Fisheries Research Murdoch University Murdoch, Western Australia, 6150, Australia Abstract — Commercial catches taken in southwestern Australian waters by trawl fisheries targeting prawns and scallops and from gillnet and longline fisheries targeting sharks were sampled at different times of the year between 2002 and 2008. This sampling yielded 33 elasmobranch species representing 17 families. Multivariate statistics elucidated the ways in which the species com- positions of elasmobranchs differed among fishing methods and provided benchmark data for detecting changes in the elasmobranch fauna in the future. Virtually all elasmobranchs caught by trawling, which consisted predominantly of rays, were discarded as bycatch, as were approximately a quarter of the elasmobranchs caught by both gillnetting and longlining. The maximum lengths and the lengths at maturity of four abundant bycatch species, Heterodontus portusjacksoni, Aptychotrema vincentiana , Squatina australis, and Myliobatis australis, were greater for females than males. The L50 determined for the males of these species at maturity by using full clasper calcification as the criterion of maturity did not differ significantly from the corresponding L50 derived by using gonadal data as the crite- rion for maturity. The proportions of the individuals of these species with lengths less than those at which 50% reach maturity were far greater in trawl samples than in gillnet and longline samples. This result was due to differences in gear selectiv- ity and to trawling being undertaken in shallow inshore waters that act as nursery areas for these species. Sound quantitative data on the spe- cies compositions of elasmobranchs caught by commercial fisheries and the biological characteristics of the main elasmobranch bycatch species are crucial for developing strategies for conserving these important spe- cies and thus the marine ecosystems of which they are part. Manuscript submitted 15 December 2009. Manuscript accepted 28 January 2010. Fish. Bull. 108:365-381 (2010). The views and opinions expressed or implied in this article are those of the author (or authors) and do not necessarily reflect the position of the National Marine Fisheries Service, NOAA. The impact of commercial fisheries on the populations of sharks and rays has, in recent years, become an issue of international concern (Stevens et al., 2000; Walker et al., 2005). It is important to recognize, however, that elasmobranchs are not only tar- geted by certain fisheries, but also comprise a substantial component of the bycatch of commercial fisheries, such as those employing trawl nets, gillnets, and longlines (Stevens et al., 2000; Stobutzki et al., 2001; Walker, 2005a). An assessment of the impacts of commercial fishing on the elasmo- branchs taken as bycatch is hindered by the fact that most of that catch is typically reported as “unidentified shark” or “mixed fish,” or not reported at all (Walker, 2005a). The numbers of sharks and rays taken as bycatch by commercial fisheries may, in some cases, exceed those of the targeted species, and many of those individuals die either during capture or after they are discarded (Bonfil, 1994; Stobutzki et al., 2002; Walker, 2005a). Although, in most studies of commercial fisher- ies, nontargeted species are referred to as bycatch, Walker et al. (2005) emphasized that some of those species are usually retained and thus consti- tute byproduct, whereas the others are usually discarded and therefore constitute bycatch in the strict sense. In an assessment of various com- mercial fisheries throughout the world, it was shown that trawl fisheries tar- geting prawns generate the largest amount of bycatch (Cook, 2003) and that the mortality of individuals in that bycatch is substantial (Bonfil, 1994). Indeed, it has been estimated that approximately two thirds of the elasmobranchs caught as bycatch in Australia’s northern prawn trawl fish- ery die while in the net (Stobutzki et al., 2002). Furthermore, that study demonstrated that most of these indi- viduals are small, and more than half are immature and some are caught immediately after birth. Many elasmobranchs are at or near the apex of marine food webs and thus their removal can have a sig- nificant impact on the trophic struc- ture of an ecosystem (Camhi et al., 1998; Stevens et al., 2000; Shepard and Myers, 2005). Furthermore, certain biological characteristics of elasmobranchs, such as their long life spans, low fecundities, and late ages at maturity, limit their ability both to withstand fishing pressure, either when targeted or caught incidentally, and to recover from overexploitation (Stevens et al., 2000; Walker, 2005a; Gallucci et al., 2006). In general, the populations of endemic species or those that have localized distribu- tions tend to be most prone to over- fishing (Stevens et al., 2000). More- over, there are little or no data on the reproductive biology of most bycatch species. Such data are required for determining the resilience of these species to fishing pressure, thereby enabling the development of manage- 366 Fishery Bulletin 108(4) ment plans for conserving their populations (Frisk et al., 2001; Stobutzki et al., 2002; Walker, 2005b). Most of the species in Australia’s rich diversity of elasmobranchs are endemic and occupy demersal habi- tats on the continental shelf or slope (Last and Ste- vens, 2009) and are thus potentially prone to depletion by demersal fishing methods such as trawling, gill- netting, and longlining (Stobutzki et al., 2001, 2002; Coelho et al., 2003; Perez and Wahrlich, 2005, Walker et al., 2005). Analyses of fisheries data for 1994 to 1999 showed that the bycatch species Heterodontus poi'tusjacksoni (Heterodontidae) and Myliobatis australis (Myliobatidae) are abundant in catches of the temperate demersal gillnet and longline fisheries of southwestern Australia (McAuley and Simpfendorfer, 2003). These two species and the rhinobatid Aptychotrema vincentia- na and the squatinid Squatina australis, which are also caught as bycatch by commercial fisheries, collectively contributed as much as 17% to the total biomass of the 172 species of fish caught during extensive trawling along the lower west coast of Australia (Hyndes et al., 1999). It has been estimated that approximately half of the elasmobranchs taken as bycatch by commercial trawlers in this region are likely to die during or after capture (Laurenson et al.1). Despite the potentially det- rimental effects of commercial fishing on the above four elasmobranch species and their ecological importance in the temperate waters of southern Australia, sound bio- logical data have been collected only for H. portusjack- soni (McLaughlin and O’Gower, 1971; Tovar-Avila et al., 2007; Jones et al., 2008; Powter and Gladstone, 2008). The families to which the above four species belong are represented elsewhere in the world by species that are taken in substantial numbers as bycatch. For ex- ample, in the eastern Pacific, up to a thousand individu- als of the heterodontid shark Heterodontus mexicanus may be caught in a single gillnet set, many of which die (Garayzar, 2006). In some parts of the world, the popu- lations of several species of rhinobatid and squatinid have been so drastically depleted from overfishing that they have been listed as critically endangered (Morey et al., 2006; Lessa and Vooren, 2007). In eastern Aus- tralia, the commercial landings of Myliobatis australis have been steadily increasing to the point where their stocks now need to be monitored (White et al., 2006). The first aim of the present study was to determine the numbers, and thereby the percent contributions, of the females and males of each shark and ray species in samples of commercial catches taken in southwestern Australian waters by demersal trawling for prawns and scallops and by demersal gillnetting and longlining for sharks. These data enabled the percent contributions made by the bycatch and byproduct to the total elasmo- 1 Laurenson, L. J. B., P. Unsworth, J. W. Penn, and R. C. J. Lenanton. 1993. The impact of trawling for saucer scallops and western king prawns on the benthic communities in coastal waters off south-western Australia, 93 p. Fisheries Res. Report, Department of Fisheries, no. 100 (part 1), Western Australia. branch catch to be estimated for each fishing method. The percent contributions of each species to the catches obtained by trawling, gillnetting, and longlining were then used to compare statistically the species composi- tions of the elasmobranchs in the catches taken by each of these fishing methods. Emphasis was next placed on determining the length compositions of the bycatch species H. portusjacksoni, A. vincentiana, S. australis , and M. australis in samples collected by trawling, gillnetting, and longlining and on estimating the lengths at maturity of the last three of these ecologically important species. The lengths of both the females and males at which 50% are mature (L50) were first determined by using gonadal data as the criterion for determining maturity and then, in the case of males, by employing full clasper calcification as that criterion. The L50 calculated for the males of each species by using the two maturity criteria were then compared to ascertain whether the L50 derived by using clasper cal- cification as the index of maturity is a reasonable proxy for that derived using gonadal stage as that criterion. The L50 at maturity of A. vincentiana, S. australis, and M. australis and of H. portusjacksoni (Jones et al., 2008) were then employed to determine the proportions of both sexes of each of these four species that were caught by each fishing method before they had the opportunity to reproduce. Finally, the management and conservation implications of our data are considered. Materials and methods Sampling regime The elasmobranchs were examined in commercial catches from 69 demersal trawls, 24 demersal gillnets, and 19 longline sets of fishing vessels operating in southwestern Australian waters southwards of 32° S lat. on the west coast and then eastwards to 118° E long, on the south coast (Fig. 1). Trawling targets the western king prawn (Melicertus latisulcatus) and the Ballot’s saucer scallop (Amusium balloti), whereas demersal gillnetting and longlining target mainly the gummy shark (Mustelus antarcticus ), the dusky shark ( Carcharhinus obscurus), and the whiskery shark ( Furgaleus macki) (McAuley and Simpfendorfer, 2003). Sampling, which was under- taken between November 2002 and November 2008, was designed to ensure that the catches of each fishing method were sampled at least once, and generally on at least three occasions, in each season of the calendar year. The trawl, gillnet, and longline fishermen, whose catches were selected for sampling, were those that fished regularly and readily allowed us onboard, and whose methods were representative of those used in the area that is fished. The numbers of each elasmobranch species caught by each fishing method on each sampling occasion were recorded, as also were the sexes of all individuals except for those of a few species in a small number of trawl samples when the catches of those spe- cies were particularly large, in which case the sexes of Jones et al.: Species compositions of elasmobranchs caught by three commercial fishing methods 367 each individual in a large, randomly selected subsample were recorded. Each elasmobranch species caught by gillnetting and longlining was categorized as either targeted, byproduct, or bycatch, whereas those taken by trawling, where prawns and scallops are the target spe- cies, were categorized as either byproduct or bycatch. Trawling was undertaken mainly at depths of 8 to 13 m in a marine embayment on the lower west coast of Western Australia and, to a lesser extent, at depths <32 m and at distances within 20 km from the main- land along that coast. The trawl net, in which the co- dend consisted of 45-mm mesh, was towed for 60-180 min at a speed of ~6.5 km/h. Commercial gillnet fisher- men deployed up to 7000 m of either 165- or 178-mm stretched mesh net that was set for up to 24 hours at depths of 24-73 m, whereas longlines, consisting of 360 hooks attached to approximately 6400 m of mainline, were set for an average of 3 hours at depths of 65-73 m. Multivariate analysis The square root of the percent contribution of the number of each elasmobranch species to the total catch of all elasmobranchs recorded in each sample during regular onboard observations of the catches taken by each fishing method was used to construct a Bray-Curtis similarity matrix, which was then subjected to nonmetric multidi- mensional scaling (nMBS). One-way analysis of simi- larities (ANOSIM) was used to test whether the species compositions of the elasmobranch catches taken by the three fishing methods were significantly different and, if so, pair-wise ANOSIM tests were used to test for dif- ferences between the compositions of the elasmobranchs obtained by each pair of methods. The R-statistic value was then employed to ascertain the extent of any differ- ences between the compositions of those catches (Clarke, 1993). R-statistic values approaching 1 demonstrate that the species composition of the a priori groups dif- fered markedly, and a value of approximately 0 indi- cates that the species compositions of those groups are very similar. Similarity percentages (SIMPER; Clarke, 1993) were used to identify the species that typify the samples obtained by each fishing method and which species are responsible for discriminating between the samples caught by each pair of methods. The ordination and associated tests were undertaken with the PRIMER vers. 6 statistical package (Clarke and Gorley, 2006). Length and weight measurements A randomly selected subset of individuals collected during regular onboard observations, together with additional randomly selected individuals provided by commercial fishermen, yielded a total of 516 H. portus- jacksoni, 340 A. vincentiana, 362 S. australis, and 218 M. australis, which were brought to the laboratory and processed. The data derived from the samples were used to construct length-frequency histograms and to derive sex ratios and reproductive data for the species. The sex of each individual was recorded and the total length 114°E 115°E 116°E 117°E 118°E 119°E Figure 1 Map showing the region between the northernmost and east- ernmost limits of southwestern Australia from which elas- mobranchs were recorded onboard commercial trawling (dark gray area) and gillnetting and longlining vessels (light and dark gray areas). (TL — tip of snout to tip of tail) and total weight (W) of each H. portusjacksoni, A. vincentiana and S. australis were measured and weighed to the nearest 1 mm and 1 g, respectively. The disc length ( DL — tip of snout to the junction of the tail and pelvic fins) of each M. aus- tralis was measured to the nearest 1 mm and the disc width (DW) and weight of each fully intact individual were recorded to the nearest 1 mm and 1 g, respectively. The relationships between DW and DL, DL and DW, and between the natural logarithms of W and DL and of W and DW for intact M. australis were calculated by using least trimmed squares (LTS) regression. A resam- pling test was used to demonstrate that the above data for the two sexes could be pooled. The above relation- ships were used to estimate the DW and W of the small number of M. australis (-15%) whose pectoral fins had been removed by fishermen for commercial sale. Note that, in the results, the sample size (n) and coefficient of determination (r2) refer to the trimmed data for the individuals used for the analyses. Length at maturity The reproductive tracts of the females and males of A. vincentiana, S. australis, and M. australis were assigned to one of the following maturity stages by using the crite- ria outlined in White et al. (2001). For females, stage 1 = uteri small and thin and oocytes not macroscopically vis- ible; stage 2 = uteri enlarging but still thin and oocytes becoming visible but not yet containing yolk; stage 3 = uteri enlarged and oocytes yolked; stage 4 = pregnant, and stage 5 = uteri or cloaca distended, indicating that parturition had recently occurred. For males, stage 1 = seminal vesicles small and thin and testes not well defined; stage 2 = seminal vesicles enlarging and start- ing to become coiled, but testes not yet lobed; stage 3 = 368 Fishery Bulletin 108(4) seminal vesicles tightly coiled and testes lobed; and stage 4 = similar to stage 3, but with semen present in the distal portion of the seminal vesicle. Individuals with reproductive tracts at stages 1 and 2 are regarded as sexually immature, whereas those at stages 3 and 4 and, in females, also at stage 5, can reproduce or have reproduced and are therefore considered mature. When we alternatively employed full clasper calcification as the indicator of maturity, we considered that males with noncalcified and partially calcified claspers were sexu- ally immature, and those with fully calcified claspers were mature because they have the ability to copulate. The probability (P) that an individual is mature was assumed to be a logistic function of its length (L): P = jl + exp[-(a + /?L)]} , (1) where a and /3 are parameters that determine the loca- tion and shape of the logistic curve. The parameters a and ft were transformed to the lengths L50 (TL50 or DL50) and L95 (TL95 or DL95) by which 50% and 95% of fish have attained maturity, respectively, using the equations ^50 ~ Otl P, (2) L95 = D°ge(19) -«]/£■ (3) The equation for the probability that a fish is mature thus becomes P = {l + exp[-loge(19)(L-L50)/(L95-L50)]} \ (4) Logistic relationships of the above form were derived for females and males of A. vincentiana, S. australis , and M. australis by using gonadal maturity status, i.e., >stage 3, as the criterion for maturity and, in addition, for males, employing full clasper calcification as the criterion for maturity. Logistic regression analysis was used to fit these logistic curves by using Solver in Excel (Microsoft, Redmond, WA) to maximize the log-likeli- hood. We used likelihood-ratio tests (Cerrato, 1990) to compare the L50 of the females and males of each spe- cies at maturity using gonadal status as the criterion for maturity and to compare the L50 derived for males using both gonadal and clasper calcification status as the criteria for maturity. For the likelihood-ratio tests, the hypothesis that the data for both the females and males of each species could be described by a common logistic curve was rejected at the a=0.05 level of signifi- cance if the test statistic, calculated as twice the dif- ference between the log-likelihoods obtained by fitting maturity curves with a common value of L50 for both sexes and by fitting separate maturity curves for each sex, exceeded ^2„(q), where q is the difference between the numbers of parameters in the two approaches. Note that the L50 of females and males of H. portusjacksoni at maturity had been determined previously with this approach (Jones et al., 2008). WinBUGS (Lunn et al., 2000) was also used to fit the logistic curves to the maturity-at-length data for the females and males of each species and to calculate the proportion that were mature at length, thereby enabling derivation of the upper and lower confidence intervals for the L50 and L95 values and for the proportion of mature males and females in each 50-mm length class (see Jones et al. [2008] for further details of this WinBUGS analysis). Results Species compositions by fishing method A total of 4820 individual elasmobranchs, representing 10 families and 22 species of sharks and 7 families and 11 species of rays, were recorded during regular onboard examinations of the catches of commercial trawl, gillnet, and longline vessels operating off the southwestern coast of Australia (Tables 1-3). The 2986 elasmobranchs caught by prawn and scal- lop trawling were dominated by rays, which comprised 10 of the 14 species and contributed 87% to the total elasmobranch catch (Table 1). The species of a single family of rays, the Urolophidae, comprising four spe- cies and two genera, contributed as much as 67% to the total trawl catch of elasmobranchs. The two species of shark ( H . portusjacksoni and S. australis) and the two species of ray (A. vincentiana and M. australis), whose biological characteristics were determined (see later), each contributed between 4.5% and 8% to the total number of elasmobranchs caught by trawling and collectively as much as 25% (Table 1). Two species of shark, M. antarcticus and C. brevipinna, which were caught in very small numbers, were retained and thus constituted byproduct. Gillnet catches yielded 1260 elasmobranchs, repre- senting 19 species of shark and 6 species of ray, with sharks contributing 96% to the total catch of elasmo- branchs (Table 2). The most abundant species in the gillnet catches was a targeted shark, Carcharhinus obscurus, which contributed more than a third to the total elasmobranch catch. The other two targeted spe- cies, Mustelus antarcticus and Furgaleus macki, which were dominated by females, ranked third and fifth in terms of abundance, respectively, and contributed an additional 14.2% and 6.5%, respectively (Table 2). The second most numerous species, however, was the shark H. portusjacksoni, a bycatch species, which constituted one-fifth of the total catch. None of the six species of ray caught by gillnetting was abundant in the catches obtained by this method. The byproduct and bycatch species contributed 18.7% and 25.7%, respectively, to the total gillnet catch. Thirteen of these species, which contributed more than one third to the total number of individuals of elasmobranchs caught, were always discarded as bycatch. The 22 species of elasmobranch caught by longlining were dominated by one of the three targeted species, M. antarcticus, which made up 63% of the total catch of 574 individual elasmobranchs (Table 3). The next four most Jones et al.: Species compositions of elasmobranchs caught by three commercial fishing methods 369 Table 1 Number of females and males and percentage contribution of females of each elasmobranch species (sex determined during regu- lar onboard observations) in the catches of trawl vessels fishing for prawns and scallops on the lower west coast of Australia. The total number of individuals of each species (including those whose sex could not be determined owing to logistic constraints) and the percent contribution of each species to the total elasmobranch catch are also given. Plain font* = byproduct species, i.e., those that are not targeted but usually retained; Bold font = bycatch species, i.e., those that are usually discarded. Common name Species name Female n Male n Female % Total n Total % Lobed stingaree Urolophus lobatus 422 288 59.4 851 28.5 Sparsely-spotted stingaree Urolophus paucimaculatus 235 349 40.2 726 24.3 Masked stingaree Trygonoptera personata 84 66 56.0 277 9.3 Western shovel nose ray Aptychotrema vincentiana 124 99 55.6 237 7.9 Port Jackson shark Heterodontus portusjacksoni 112 110 50.5 223 7.5 Southern fiddler ray Trygonorrhina dumerilii 94 118 44.3 220 7.4 Western shovelnose stingaree Trygonoptera mucosa 23 27 46.0 153 5.1 Southern eagle ray Myliobatis australis 67 81 45.3 148 5.0 Australian angelshark Squatina australis 67 66 50.4 135 4.5 Smooth stingray Dasyatis brevicaudata 2 6 25.0 8 0.3 Gummy shark* Mustelus antarcticus* 4 1 80.0 5 0.2 White-spotted guitarfish Rhynchobatus australiae 1 0 100.0 1 < 0.1 Spinner shark* Carcharhinus brevipinna * 0 1 0 1 < 0.1 Coffin ray Hypnos monopterygius 0 1 0 1 < 0.1 Total 1235 1213 2986 abundant species, which were all bycatch, consisted of three species of rays and the shark H. portusjacksoni and collectively contributed nearly a quarter of the individual elasmobranchs obtained by longlining. The other two targeted species, C. obscurus and F. macki, contributed only 3.8% and 0.9%, respectively, to the total catch taken by longlining. The 19 nontargeted spe- cies caught by longlining comprised eight species that were always discarded as bycatch and represented one quarter of the total catch. On the ordination plot, derived from the similarity matrix constructed by using percent contributions of the various species to the elasmobranch catches obtained by the three fishing methods, the samples for longlin- ing lie above those for gillnetting, and both of these lie to the right of the discrete group comprising the trawl samples (Fig. 2). One-way ANOSIM confirmed that the compositions of the samples obtained by the three fishing methods were significantly different (P=0.001, global R = 0.753). Pair-wise ANOSIM tests revealed that the compositions in the samples collected by each method differed significantly from those in the samples obtained by each other method (all PcO.OOl), and the R- statistic was similarly high for trawling vs. gillnetting (0.797) and trawling vs longlining (0.774) and greater than that for gillnetting vs longlining (0.515). The most important of the typifying species for the trawl samples, i.e., those that were most abundant and were found most frequently, comprised two ray spe- cies, A. vincentiana and Urolophus paucimaculatus, and two shark species, H. portusjacksoni and S. aus- tralis (Table 4). Heterodontus portusjacksoni and M. Figure 2 Nonmetric multidimensional scaling (nMDS) ordina- tion plot, derived from the matrix constructed from the percentage contribution (square root transformed) of each elasmobranch species recorded during each of the regular onboard observations of the catches of com- mercial trawling (open circles), gillnetting (gray circles), and longlining vessels (black circles). antarcticus were also important typifying species for the elasmobranch catches taken by both gillnetting and longlining. Carcharhinus obscurus is a particularly important typifying species for the gillnet samples and the same is true for the bycatch ray species Dasyatis brevicaudata for the longline samples. Relatively greater 370 Fishery Bulletin 108(4) Table 2 Number of females and males, percent contribution of females, and the total number and percent contributions of each elasmo- branch species that were recorded during regular onboard observations of catches from gillnet vessels on the southwest coast of Australia. Plain font = targeted species; Plain font* = byproduct species i.e., those species not targeted but usually retained; Bold font = bycatch species, i.e., those species that are usually discarded. Common name Species name Female n Male n Female % Total n Total % Dusky shark Carcharhinus obscurus 229 208 52.4 437 34.7 Port Jackson shark Heterodontus portusjacksoni 131 120 52.2 251 19.9 Gummy shark Mustelus antarcticus 151 28 84.4 179 14.2 Sandbar shark* Carcharhinus plumbeus* 86 59 59.3 145 11.5 Whiskery shark Furgaleus macki 69 13 82.9 82 6.5 Southern eagle ray Myliobatis australis 18 14 56.3 32 2.5 Spinner shark* Carcharhinus brevipinna* 17 8 68.0 25 2.0 Western wobbegong* Orectolobus hutchinsi* 7 11 38.9 18 1.4 Gulf wobbegong* Orectolobus halei* 3 12 20.0 15 1.2 Smooth hammerhead* Sphyrna zygaena* 8 4 66.7 12 1.0 Cobbler wobbegong Sutorectus tentaculatus 7 4 63.6 11 0.9 Bronze whaler* Carcharhinus brachyurus * 10 1 90.9 11 0.9 Spotted wobbegong* Orectolobus maculatus* 2 7 22.2 9 0.7 Western shovelnose ray Aptychotrema vincentiana 2 4 33.3 6 0.5 Australian angelshark Squatina australis 2 4 33.3 6 0.5 Common sawshark Pristiophorus cirratus 4 1 80.0 5 0.4 Southern fiddler ray Trygonorrhina dumerilii 3 2 60.0 5 0.4 Grey nurse shark Carcharias taurus 2 0 100.0 2 0.2 Lobed stingaree Urolophus lobatus 1 1 50.0 2 0.2 Floral banded wobbegong Orectolobus floridus 2 0 100.0 2 0.2 Smooth stingray Dasyatis brevicaudata 1 0 100.0 1 < 0.1 Pencil shark* Hypogaleus hyugaensis* 0 1 0.0 1 <0.1 Ornate angelshark Squatina tergocellata 1 0 100.0 1 <0.1 Scalloped hammerhead* Sphyrna lewini* 0 1 0.0 1 <0.1 Western shovelnose stingaree Trygonoptera mucosa 0 1 0.0 1 <0.1 Total 756 504 1260 and more consistent numbers of A. vincentiana were particularly important for discriminating between the compositions of the samples caught by trawling and those obtained by both gillnetting and longlining, and greater and more consistent numbers of C. obscurus were especially important for discriminating between the samples taken by gillnetting from those obtained by both trawling and longlining (Table 4). The longline samples were discriminated from those obtained by both trawling and gillnetting by consistently greater numbers of D. brevicaudata. Length-frequency compositions of the four selected bycatch species by fishing method Wide size ranges of H. portusjacksoni, A. vincentiana , and S. australis and, to a certain extent, M. australis, were caught by trawling. However, the lengths of most H. portusjacksoni and M. australis were small and thus lay toward the lower end of their length ranges (Fig. 3). Although gillnetting also caught a broad size range of both H. portusjacksoni and A. vincentiana, it yielded predominantly larger S. australis and medium-size M. australis (Fig. 3). Although longline catches contained a wide size range of H. portusjacksoni and M. australis, they did not include the smallest individuals of these two species and only one of the A. vincentiana caught by this method was small (Fig. 3). No S. australis was caught by longlining. The H. portusjacksoni obtained by all three fishing methods ranged from 180 to 1300 mm TL (Table 5), the latter length rarely being exceeded by this species throughout its range (Last and Stevens, 2009). The smallest individuals possessed conspicuous um- bilical scars and were therefore neonates. The length- frequency distribution of female H. portusjacksoni is trimodal, whereas that of males is bimodal, and these modes correspond to the first two modes of females (Fig. 4). These differences account for the lengths of many females greatly exceeding the maximum length of 815 mm for males (Table 5). The weights of H. por- tusjacksoni ranged from 39 to 12,250 g (Table 5). The Jones et al.: Species compositions of elasmobranchs caught by three commercial fishing methods 371 TabSe 3 Number of females and males, percent contribution of females and the total number and percent contributions of all individuals of each elasmobranch species that were recorded during regular onboard observations of the catches from longline vessels on the southwest coast of Australia. Plain font = targeted species; Plain font* = byproduct species, i.e., those that are not targeted but usually retained; Bold font = bycatch species, i.e., those species that are usually discarded. Common name Species name Female n Male n Female % Total n Total % Gummy shark Mustelus antarcticus 234 129 64.5 363 63.2 Smooth stingray Dasyatis brevicaudata 21 20 51.2 41 7.1 Southern eagle ray Myliobatis australis 15 19 44.1 34 5.9 Southern fiddler ray Trygonorrhina dumerilii 24 8 75.0 32 5.6 Port Jackson shark Heterodontus portusjacksoni 20 11 64.5 31 5.4 Dusky shark Carcharhinus obscurus 16 6 72.7 22 3.8 Smooth hammerhead* Sphyrna zygaena* 8 1 88.9 9 1.6 Bronze whaler* Carcharhinus brachyurus * 4 1 80.0 5 0.9 Whiskery shark Furgaleus macki 5 0 100.0 5 0.9 Western wobbegong* Orectolobus hutchinsi * 1 4 20.0 5 0.9 Common sawshark* Pristiophorus cirratus * 4 1 80.0 5 0.9 School shark* Galeorhinus galeus* 2 2 50.0 4 0.7 Western shovelnose ray Aptychotrema vincentiana 3 0 100.0 3 0.5 Gulf wobbegong* Orectolobus halei* 1 2 33.3 3 0.5 Sandbar shark* Carcharhinus plumbeus* 2 0 100.0 2 0.4 Spotted wobbegong* Orectolobus maculatus * 0 2 0.0 2 0.4 Rusty carpetshark Parascyllium ferrugineum 0 2 0.0 2 0.4 Melbourne skate Spinirija whitleyi 2 0 100.0 2 0.4 Spinner shark* Carcharhinus brevipinna* 0 1 0.0 1 0.2 Australian sawtail catshark Figaro boardmani 1 0 100.0 1 0.2 Pencil shark* Hypogaleus hyugaensis* 0 1 0.0 1 0.2 Scalloped hammerhead* Sphyrna lewini* 1 0 100.0 1 0.2 Total 364 210 574 ratio of females to males of H. portusjacksoni differed significantly from parity among all individuals collec- tively (1 female:0.76 males; j2=9.46, PcO.Ol), but not for juveniles (1 female:1.20 males; ^2=2.65, P>0.05). Note that, when calculating the sex ratios for juveniles, the term juvenile refers to females and males with lengths less than the smallest mature individual of their re- spective sex. The A. vincentiana caught by all three fishing meth- ods ranged from 201 to 1001 mm TL (Fig. 4; Table 5), and the smallest individuals lay within the length range recorded for the embryos of this species (A. Jones, unpubl. data) and the largest individuals exceeded the length of “at least 840 mm” reported for this species by Last and Stevens (2009). The length-frequency dis- tributions of female and male A. vincentiana were both broadly bimodal and the numbers of both sexes were rel- atively low, between 500 and 699 mm (Fig. 4). However, the modal length class of 850-899 mm for the group of large females far exceeded that of 700-749 mm for the group of large males. Furthermore, the largest female A. vincentiana was both far longer (1001 mm) and heavi- er (3634 g) than the largest male, i.e., 872 mm and 1886 g, respectively (Table 5). The ratio of females to males differed significantly from parity among all individu- als (1 female:0.67 males; ^2=12.81, PcO.001), but not among juveniles (1 female:0.76 males; ^2=3.52, P>0.05). The smallest S. australis caught by all three fishing methods was 228 mm in TL (Fig. 4; Table 5) and thus only slightly longer than the length of 220 mm recorded for the largest embryo of this species in a concomitant study (A. Jones, unpubl. data). Although the maximum length of 1004 mm for S. australis in our samples is considerably less than the maximum length reported for this species by Last and Stevens (2009), it is still far greater than the TL50 for either females or males at maturity in southwestern Australian waters. Although individuals were represented in all 50-mm length class- es between 200 and 1049 mm, the length-frequency distributions of females and males were both dominated by their 250-299 mm length classes (Fig. 4). The larg- est female S. australis was far longer (1004 mm) and heavier (10,970 g) than the largest male (859 mm and 5500 g) (Table 5). The ratio of females to males of S. australis did not differ significantly from parity among either all individuals collectively (1 female:1.05 males; j2 = 0.18, P>0.05) or among juveniles (1 female:l.ll males; j2=0.75, P>0.05). 372 Fishery Bulletin 108(4) Table 4 Main species that typified the catches of elasmobranchs recorded onboard trawl, gillnet, and longline vessels (shaded back- ground), and those that discriminate between the catches of elasmobranchs obtained by each pair of fishing methods (unshaded background). Plain font = targeted species; Bold font = bycatch species; * Denotes that the species is relatively more abundant and consistently caught by the sampling method on the top (horizontal) row than on the side (vertical) column. Trawl Gillnet Longline Trawl Aptychotrema vincentiana Heterodontus portusjacksoni Urolophus paucimaculatus Squatina australis Gillnet Carcharhinus obscurus Aptychotrema vincentiana* Carcharhinus obscurus Heterodontus portusjacksoni Heterodontus portusjacksoni Mustelus antarcticus Mustelus antarcticus Myliobatis australis* Furgaleus macki Furgaleus macki Longline Dasyatis brevicaudata Carcharhinus obscurus* Aptychotrema vincentiana * Dasyatis brevicaudata Dasyatis brevicaudata T rygon orrh in a dumerilii* Mustelus antarcticus Mustelus antarcticus Heterodontus portusjacksoni* Heterodontus portusjacksoni* Trygonorrhina dumerilii Furgaleus macki* Heterodontus portusjacksoni Table 5 Biological characteristics of four elasmobranch species caught as bycatch by commercial trawl, gillnet, and longline fisheries operating off southwestern Australia. Length measurements are given as total lengths (TL) for Heterodontus portusjacksoni, Aptychotrema vincentiana , and Squatina australis, and as disc lengths ( DL ) for Myliobatis australis. * denotes a value extrapo- lated from the regression equation of the relation between DL and W. The true weight could not be recorded because the pectoral fins had been removed by fishermen. Sample size for each sex of each species is shown on Figure 4. Heterodontus portusjacksoni Aptychotrema vincentiana Squatina australis Myliobatis australis Females Length range (mm) 198-1300 201-1001 228-1004 118-800 Weight range (g) 39-12,250 32-3634 94-10,970 117-37,811* Smallest mature (mm) 715 754 825 444 Largest immature (mm) 869 895 834 472 Males Length range (mm) 180-815 214-872 246-859 129-545 Weight range (g) 39-3920 33-1886 115-5500 152-12,373* Smallest mature (mm) 595 642 754 365 Largest immature (mm) 654 792 707 433 The relationships between DW and DL for females and males of M. australis collectively are described by the following equations: DW = 1.70 DL + 8.16 (r 2=0.997, n = 96), (5) DL = 0.58 DW - 3.01 (r 2=0.998, n = 96). (6) From the values obtained from the above equations, the following relationships between the natural loga- rithms of W and DL, and of W and DW for both sexes are described as loge W = 2.91 loge DL - 8.92 (r 2=0.999, n=91), (7) loge W = 3.19 loge DW - 12.25 (r 2=0.999, n= 91). (8) After correction for bias (Beauchamp and Olson, 1973), the respective back-transformed relationships became Jones et al.: Species compositions of elasmobranchs caught by three commercial fishing methods 373 B Gillnet 0 200 400 600 800 1000 1200 1400 Length (mm) Figure 3 Length-frequency distributions for all males and females of Heterodontus portusjacksoni, Aptychotrema viticentiana, Squatina australis, and Myliobatis australis obtained from the commercial catches of (A) trawl, (B) gillnet, and (C) longline vessels. Length refers to total length ( TL ), except in the case of M. australis where it refers to disc length (DL). W = 0.0001344 DL2 91, (9) W = 0.000004787 DW3 19 (10) The M. australis caught by using the three sampling methods ranged from 118 to 800 mm DL (Fig. 4, Table 5), which corresponds to 198-1192 mm DW. The minimum DW is at the extreme lower end of the range reported for this species at birth, and the maximum DW is appreciably less than the maximum DW of 1600 mm recorded for M. australis (Last and Stevens, 2009). The largest female caught was 800 mm DL and 37,811 g W, which greatly exceeded the 545 mm DL and 12,373 g W of the largest male (Table 5). Both sexes were represented in each DL class between 100 and 549 mm, and females were also present in each subsequent size class up to 800-849 mm (Fig. 4). The length-fre- quency distributions for both sexes contained a single prominent modal length class at 150-199 mm. The ratio of 1 female:1.27 males of M. australis among the individuals caught by all methods collectively did not differ significantly from parity (^2=3.10, P>0.05), and the same was true for juveniles, i.e., 1 female:1.04 males (^2=0.06, P>0.05). Lengths of females and males at maturity The smallest female and male of H. portusjack- soni with mature gonads measured 715 and 595 mm TL, respectively (Table 5). Using gonadal stage as the index, we found that the TL50 for female H. portusjacksoni at maturity was 805 mm, and the TL50 for males was 593 mm, which represent 62% and 73% of their respective maxi- mum TL. The TL50 for males was only 12 mm greater and not significantly different from the 581 mm derived by using full clasper calcifica- tion as the index of maturity (Table 6, see also Jones et al., 2008). The smallest female and male of A. vin- centiana with mature gonads were 754 and 642 mm, respectively, and all females >896 mm and males >793 mm were mature (Fig. 5, Table 5). The TL50 for females of 798 mm at maturity differed significantly (P<0.001) from the corresponding TL50 for males of 671 mm when using gonadal stage as the criterion for maturity (Table 6). The latter TLg0 for males did not differ sig- nificantly from the TL50 of 654 mm derived by using full clasper calcification as the criterion for maturity (P> 0.05) (Fig. 5, Table 6). The TL50 for female and male A. vincentiana at maturity, using gonadal status as the criterion for maturity, were 80% and 77% of their respective maximum TL. On the basis of gonadal data, the TL of the smallest mature female and male S. australis were 825 and 754 mm, respectively, and all females >840 mm and all males >754 mm were mature (Fig. 5, Table 5). The TL50 for females (823 mm) and males of S. australis (734 mm), derived by employing gonadal stage as the crite- rion for maturity, were significantly different (PcO.001; Table 6). The latter TL50 was not significantly different (P> 0.05) from the TL50 of 721 mm derived for male S. australis when using clasper calcification as the crite- rion for maturity (Fig. 5, Table 6). The TL50 calculated for females and males of S. australis, with gonadal stage as the criterion for maturity, were 82% and 85% of their respective maximum TL. 374 Fishery Bulletin 108(4) The DL of the smallest females and males of M. aus- tralis with mature gonads were 444 and 365 mm, re- spectively, and all females and males with DL >513 and 433 mm, respectively, were mature (Fig. 5, Table 5). The DL50 of 511mm (DW50 = 879 mm) for females at maturity differed significantly (P<0.001) from the 399 mm (Z)W50 = 689 mm) of males when gonadal status was used as the criterion for maturity (Table 6). The latter DL50 did not differ significantly (P>0.05) from the 388 mm (DW50 = 670 mm) derived for males at maturity with clasper calcification as the criterion for maturity (Table 6). On the basis of gonadal criteria, the DL50 for females and males of M. australis at maturity were 64% and 73% of their DLmax, respectively. 40 r B Aptychotrema vinceritiana D Myliobatis australis females n=203 males a=137 females n=96 males n=122 i . i 0 200 400 600 800 1000 1200 1400 Length (mm) Figure 4 Length-frequency distributions for females (white histograms) and males (black histograms) of (A) Heterodontus portusjacksoni, (B) Aptychotrema vinceritiana, (C) Squatina australis , and (D) Myliobatis australis obtained collectively by all three fishing methods. Length refers to total length ( TL ), except in the case of M. australis where it refers to disc length (DL). Percentage of females and males caught by each fishing method The percentage of females of H. portusjacksoni, A. vincentiana, S. australis and M. australis caught in trawls with lengths below the L50 (TL5 0 or DL50) at maturity were very high and similar to those of males (Table 7). The percentage in trawl samples of both sexes with lengths less than their L50 at maturity ranged from 63% for A. vincen- tiana, to 90% for both M. australis and S. aus- tralis, respectively, to 97% for H. portusjacksoni (Table 7). In the case of gillnet samples for three of the four species, the percentage of females with lengths less than their L50 at maturity exceeded those of males and particularly so for M. australis, for which the values were 86% and 40%, respec- tively (Table 7). In longline samples, the percent- age of males of H. portusjacksoni with lengths below their L50 at maturity slightly exceeded the corresponding value for females (37% and 32%, respectively). Discussion This study is the first to quantify the contribu- tion of each elasmobranch species to the total elasmobranch catch obtained by co-occurring trawl, gillnet, and longline fisheries, and to cal- culate the contributions made by the bycatch and byproduct species to the catches taken by each fishing method. In addition, our results indicate that nMDS ordination and associated tests would be invaluable for detecting whether the species composition of elasmobranchs in the catch produced by each fishing method changes in the future in response to either variations in fishing activity or environmental factors and, if so, also for elucidating the magnitude of that effect. This study has also produced, for four abundant bycatch species, the sound quantitative biological data of the types required by manag- ers for developing plans for conserving stocks and which are deficient for the vast majority of bycatch species (Stobutzki et al., 2002). The sizes at maturity that were determined for the four bycatch species in this study enabled the proportion of each species, which was caught by each fishing method before it had the potential to reproduce, to be estimated. Jones et al.: Species compositions of elasmobranchs caught by three commercial fishing methods 375 Table 6 Estimates of the L50 and L95 at maturity, and the upper and lower 95% confidence limits (CL), for females and males of Heter- odontus portusjacksoni, Aptychotrema vincentiana, and Squatina australis recorded as total length (TL). For Myliobatis austra- lis, these values were recorded as disc length ( DL ); extrapolated values for disc widths (DW) are also provided. Estimates were derived by using gonadal status as an index of maturity for females and males and also by using full clasper calcification as that index for males. Sample sizes for females and males of each species are provided in Figure 4. Female (gonads) Male (gonads) Male (claspers) ^50 ■^95 ^50 ■^95 ■^50 •^95 Heterodontus portusjacksoni Estimate 805 896 593 647 581 652 Upper 95% CL 826 931 605 674 594 689 Lower 95% CL 781 866 579 628 563 629 Aptychotrema vincentiana Estimate 798 877 671 766 654 707 Upper 95% CL 815 920 695 833 676 756 Lower 95% CL 774 855 631 736 618 679 Squatina australis Estimate 823 852 734 735 721 723 Upper 95% CL 842 927 753 806 753 806 Lower 95% CL 771 826 673 714 674 714 Myliobatis australis Estimate 511 585 399 472 388 453 Upper 95% CL 558 696 416 518 404 491 Lower 95% CL 480 538 376 440 366 420 Female (gonads) Male (gonads) Male (claspers) dw50 dw95 dw50 dw95 dw50 dw95 Myliobatis australis Estimate 879 1006 689 813 670 781 Upper 95% CL 960 1195 717 891 697 845 Lower 95% CL 827 926 649 758 632 724 Table 7 Numbers of females and males and total number of Heterodontus portusjacksoni , Aptychotrema vincentiana, Squatina australis, and Myliobatis australis that were caught by each fishing method and examined in the laboratory, and the percentages of indi- viduals with lengths less than their L50 at maturity when using gonadal status as the index of maturity. L50 refers to total length (TL50), except in the case of M. australis where it refers to disc length ( DL50 ). Fishing method Species name Females Males Sexes combined n % < L50 n % < L50 n % < L50 Trawl Heterodontus portusjacksoni 121 98 128 96 249 97 Aptychotrema vincentiana 182 62 116 66 298 63 Squatina australis 155 89 169 91 324 90 Myliobatis australis 71 92 83 87 154 90 Gillnet Heterodontus portusjacksoni 115 44 76 30 191 38 Aptychotrema vincentiana 16 38 20 30 36 33 Squatina australis 22 18 16 19 38 18 Myliobatis australis 14 86 25 40 39 56 Longline Heterodontus portusjacksoni 57 32 19 37 76 33 Aptychotrema vincentiana 5 0 1 100 6 17 Squatina australis — — — — — — Myliobatis australis 11 36 14 7 25 20 376 Fishery Bulletin 108(4) CD -Q O "S, < L O o o CO o ea o o o CO (%) Aouenbejj o o o o o o o oo o o CD o o -Cf o o CM o o CD o o o o CNJ o o o o o CO o o CD o o •'T o o CNI o> c 0 o LO c n d CD o3 a 3 in >-d J5 ;> CO w £ lac a .g _ c/) g 3 - ^ v> Xi £ 2 w ^ o o - 03 03 -o o Oh T3 Ch 03 a 6 w CD £ ^ .O +-> Ct-i T3 | 0 S 0 c^.a \ — co '■a „ 7 0 S3 42 3 ^ be 0 .S >§ Si CD o- 3 3 W +J . 03 03 a cc ■ — a .0 < t to 0 -3 £ 03 H o £ -a . -S S >> § (O u o3 2 +J V Sh £ ,pH o be o o CD X 4-» ^ 0 g > ^ o w x _d 03 73 rH lac g S £ cd O '-H X o co .Cfl ^ 7 a -rH Oh 0.05). Growth in length and mass of Pacific cod larvae was significantly affected by rearing temperatures across the range examined (Fig. 1; gM F(311| = 30.0, P<0.001; gL (F|3 h]= 59.4, PcO.001). After 35 days, fish reared at 8°C were 2.6 times larger than fish reared at 2°C (MD 0.271 vs. 0.104 mg). Growth rates were fitted as a sec- ond-order function of temperature (Table 2). Postflexion larvae The sorting of fish by size before the establishment of experimental groups produced significant differences in initial sizes of experimental fish (group mean Ls 14.02 to 16.10 mm; P- 0.004). Differences among size groups within temperature treatment were generally maintained throughout the experiment but growth rates were slightly higher in the small-size groups than the in large-size groups. The effect of size group was significant for growth expressed as gM (P|26j=15.7, P=0.004) but not for gL (F[2 6] = 1.78, P=0.248). Growth in length and mass of postflexion Pacific cod larvae was significantly affected by rearing tempera- tures across the range examined (Fig. 2; gM Fj3 g]=34.3, P<0.001; gL (P[36j = 63.0, P<0.001). Growth rates (for gM and gL) at 11°C averaged 2.2 and 3.0 times, respec- tively those observed at 3°C in similar size treatments. Growth rates were described as a second-order function of temperature (Table 2). Juveniles Sorting of fish by size before the experiment resulted in significant differences in initial size among repli- cates within temperature treatments (LT P[8 83) = 9.46, 386 Fishery Bulletin 108(4) Temperature (°C) Figure t Temperature-dependent growth rates of preflexion larval Pacific cod ( Gadus macrocephalus ) in (A) stan- dard length and (B) dry mass. Fish used in experiments were the offspring of spawning adults collected in the central Gulf of Alaska in April 2007 (open symbols) and 2008 (filled symbols). Values are the mean growth rates (±standard error) for three replicate tanks at four temperatures. Overlapping points are displaced hori- zontally for clarity. 0.6 ■O E £ 6> 0.5 0.4 O) c embryo allometric scaling, wherein growth rates after hatching are lower than those predicted by body size allometry. We determined parameters for temperature-dependent growth functions in length and mass for specific life stages and for a unified STDG function. These measures of growth potential can be used to evaluate the biotic and abiotic factors regulating the growth and survival of early life stages of Pacific cod in the wild (Folkvord, 2005; Hurst et al., 2010). and (0.454 + 1.610 T- 0.069 T2)- g-6.725 Md _j_ 3 705 if stage = embryo 8m = ' (4) (0.454 + 1.610 T - 0.069 T2)- g-6.725 md if stage > embryo These equations explained over 88% of the observed variance in growth rates of Pacific cod embryos to post- settlement juveniles (gM r=0.969; gL r=0.940). Analysis of residuals from these models indicated greater vari- ance at higher growth rates (small body sizes and higher temperatures), but there were no trends in residuals in relation to experimental temperature or body size that would indicate significant departures from the model. Discussion Temperature is the dominant regulator of growth in early life stages of fishes. In this study we examined the ontogenic pattern in growth rate for early life stages of Pacific cod. We demonstrated a deviation from strict Experiments Because these experiments were conducted across a range of life stages, many aspects of our experimental method had to be adapted for each specific experiment, such as tank volume, prey type, and fish density. How- ever, the most significant differences in method between egg-larval and juvenile experiments were the level of observation and source of fish. To estimate growth rates of embryos and larvae, subsamples of fish were drawn from a large tank population to determine mean size at a specific age (Otterlei et al., 1999; Monk et al., 2008). Change in mean size at age was then used to determine growth rate in each replicate tank. With this approach, there is the potential for size-selective mortality in the experiments to affect estimates of mean growth rates. Such size-selective mortality is most commonly assumed to be the result of predation (including cannibalism in single species culture). However, such an effect is unlikely to have occurred in these experiments: postflex- ion larvae were sorted by size before growth experiments specifically to avoid potential cannibalism, and we saw no evidence of cannibalism in the experiment (no larvae in samples with fish in their stomachs). Because juvenile fish could be handled, they were measured and returned to the tank and subsequently remeasured, providing 388 Fishery Bulletin 108(4) °-6 I A O 0.1 T 0.0 I T T T ^ 2 4 6 8 10 Figure 3 Temperature-dependent growth rates of postsettlement juvenile Pacific cod (Gadus macrocephalus) in (A) total length and (B) wet mass. Fish used in the experiment were collected from nearshore nursery areas at Kodiak Island, Alaska in July 2008. Values are the mean growth rates (± standard error) for three replicate tanks at four temperatures. Symbols represent groups based on initial size sorting of fish (circle: small; triangle: medium; square: large). Overlapping points are displaced horizontally for clarity. growth trajectories for individual fish. These individual growth rates were then averaged to estimate mean growth rates of fish in each tank (Hurst and Abookire, 2006; Wijekoon et ah, 2009). For experiments with eggs and larvae, we used the offspring of field-caught spawning adults. In each year of experiments, gametes from one or two females were mixed with those of three to five males. Juvenile fish used in experiments were naturally produced and cap- tured after recruitment to juvenile nearshore habitats. Therefore, the genetic diversity among experimental ju- veniles was significantly greater than that found in ex- perimental eggs and larvae. This difference in potential genetic contributions to growth rates is not expected to have a significant influence on the overall growth pat- terns described here because maternal and genetic ef- fects on growth have been shown to be small in relation to environmental factors such as temperature (Benoit and Pepin, 1999; Green and McCormick, 2005) and prey availability (Clemmesen et al., 2003). Ontogenetic patterns of growth By combining results across life stages, we clarified ontogenetic patterns in growth and length. With the exception of a departure during the preflexion stage, growth patterns followed expected size-dependent allo- metric patterns throughout much of the early devel- opment (Elliott and Hurley, 1995). Growth in mass (gM) decreased with increases in body size (Wootton, 1990) and growth in length (gL) was constant across life stages. Although the constancy of gL across a range of body sizes in early life stages has been noted in other studies (Jones, 2002; Sigourney et al., 2008), those experiments have not generally included the presettle- ment egg and larval stages incorporated here. Interest- ingly, in Pacific cod, the general pattern of across-stage similarity in gL extended from the egg-embryo stage into the juvenile stage. There was a significant departure from the expect- ed ontogenetic pattern in the preflexion larval period, most clearly observable in the gL results. Measured gL among preflexion larvae averaged only 41% of the rates measured for embryos, postflexion larvae, and settled juveniles. Although less readily apparent, a similar de- parture was observed in mass growth as the decline in gM between the egg-embryo and preflexion larval stages was greater than expected from allometric patterns accounting for the observed differences in body size. In several studies on the growth of first-feeding gadids, higher growth rates were reported for fish reared on copepods than on cultured rotifers (Conceicao et al., 2010). Unfortunately, technical limitations preclude rearing sufficient quantities of copepods for use in ex- periments such as ours. For our study, larval Pacific cod were reared on essential-fatty-acid-enriched rotifers, as the best of the practicable prey alternatives. Therefore, it is possible that growth rates of preflexion larvae are under-estimates of maximum potential growth at this stage. However, the effect of prey type is insufficient to completely explain the significantly lower growth rate observed at this stage when compared to other stages. Further, similar observations of reduced growth rates of fishes in the early posthatch phase have been observed in several other studies. Experiments in which growth of haddock ( Melanogrammus aeglefinus; Martell et al., 2005) was tracked through the egg-larva transition revealed a similar reduction in growth associated with hatching. A similar pattern is apparent in Atlantic cod, but without measurements of embryonic growth rates, the magnitude of decline at hatching could not be determined (Otterlei et al., 1999; Folkvord, 2005). However, these studies document a period of increas- ing growth rates after hatching, followed by growth rate declines along allometric expectations, indicating a similar overall pattern. Hurst et al. : Growth rates of Gadus macrocephalus 389 Figure 4 Size- and temperature-dependent growth rates of early life stages of Pacific cod (Gadus macrocephalus) from the central Gulf of Alaska in (A) standard length and (B) dry mass. Points are the tank mean growth rates measured in laboratory experiments. Surfaces are best-fit descriptions of growth rates. Parameter values are presented in Equations 3 and 4. This reduction in growth during the egg-larva tran- sition appears to be the result of increased metabolic expenditures associated with swimming in posthatch larvae and possibly a reduction in energy available for growth associated with the transition from reliance on endogenous energy stores to exogenous feeding (Torres et al., 1996; Yufera and Darias, 2007). In Pacific cod, yolk reserves are depleted 3-12 dah, depending on water temperature (Laurel et al., 2008) and stomach fullness increased through the first 28 dah (B. J. Lau- rel, unpubl. data). The increase in growth rates after the preflexion transitional feeding stage coincides with the onset of diel vertical migrations (Hurst et al., 2009) and increased responsiveness to prey (Colton and Hurst, 2010) among postflexion Pacific cod larvae. The negative departure from an allometrically defined growth pattern after hatching indicates that the first- feeding stage represents a “critical period” in the early life history of Pacific cod and that the consequences for recruitment of this low growth may be greater at low temperatures (Kamler, 1992; Houde, 1996). In addition to inclusion of embryo measurements into larval studies, future studies with other species should encompass other major life history and habitat transi- tions, such as metamorphosis and settlement in flat- fishes (Christensen and Korsgaard, 1999; Neuman et al., 2001) in order to clarify the physiological basis of growth patterns and to determine parameters for growth models. Based on exploration of model structures for a unified STDG for early life stages of Pacific cod, a two-stage model was developed. The first stage described growth in the egg stage as a direct function of water tempera- ture. The second stage described growth in posthatch fish as a function of water temperature and fish size. In addition to providing the best fit to experimental data, this formulation is logically consistent with the life history. Explicit discrimination between life stages coincides with hatching, whereas the function provides a continuous growth surface for all free-swimming stag- es. This stage-independent model for posthatch fish provides more realistic growth trajectories in modeling applications where fish are tracked over multiple stages. Applications of growth models By quantitatively accounting for the influence of tem- perature variation, laboratory-determined growth rates are being increasingly used to evaluate factors regulat- ing growth rates of fishes in the wild (Folkvord, 2005; Rakocinski et al., 2006; Hurst et al., 2009). In these analyses, “realized growth” expresses observed growth in the field as a fraction of the potential growth at the encountered field temperatures (Hurst and Abookire, 2006), with field growth rates estimated from changes in mean size, otolith increment measures, or biochemi- cal measures (RNA:DNA). In these studies it is impor- tant to recognize that growth rates of individuals are determined by both genetic and environmental factors. Growth models from laboratory experiments generally describe mean growth of a representative population under optimal foraging conditions, which should not be mistaken for the maximum growth rates that would be observed for the fastest growing individual. Therefore, field growth rates should be similarly expressed as a population mean rather than at the individual level. Realized growth rates near 100% indicate that growth rates in the population are directly limited by ambient 390 Fishery Bulletin 108(4) temperature variation. In studies of juvenile flatfishes, this temperature regulation of growth has been referred to as the “maximum growth/optimal food condition” hypothesis (Karikiri et al., 1991; van der Veer and Witte, 1993). Conversely, realized growth rates signifi- cantly below 100% indicate that growth is regulated by nonthermal environmental factors such as light regime or prey availability (Buckley et al. 2006; Kristiansen et al., 2007; Hurst et al., 2009). Unfortunately, many studies of growth rates in fishes are conducted over a limited size range and usually within a single life stage. Therefore, these data have limited application where growth rates of wild fish are tracked over longer time periods or through early life history stage transitions. For example, in evaluating the mechanisms responsible for variation in survival and recruitment, it is critical to determine whether growth reductions among wild fish are due to inher- ent physiologically based patterns (as appears in post- hatch gadids) or are imposed by an unfavorable growth environment (Jones, 2002). In another application of laboratory data to field studies, the back-calculation of hatch dates from estimated temperature-dependent growth rates (Lanksbury et al., 2007) could be biased if ontogenetic patterns in growth variation are not ac- counted for. Conclusion Growth variation in early life stages can result in body- size variation that persists over time and has signifi- cant implications for the survival and recruitment of marine fish larvae (Houde, 1996; Jones, 2002). Success- ful evaluation of the biotic and abiotic factors regulating this underlying variation in growth requires detailed information on the size- and temperature-dependency of potential growth throughout the early life history. Identifying the intrinsic patterns in growth-rate allom- etry and reductions among preflexion larval Pacific cod was based on the integration of experimental data on embryos and larvae, — stages generally considered in isolation from each other. We suggest that data on embryos be routinely incorporated with larval data to clarify ontogenetic and temperature-dependent growth patterns in the early life history stages of fish. Acknowledgments We thank T. Tripp, M. Spencer, and B. Knoth for assis- tance with fish collection and shipping. Staff and stu- dents in the Fisheries Behavioral Ecology Program, including L. Copeman, S. Haines, M. Ottmar, P. Iseri, A. Colton, L. Logers, E. Seale, and J. Scheingross assisted with various laboratory experiments. S. Munch, L. Tomaro, A. Stoner, and three anonymous reviewers provided helpful comments on earlier versions of this manuscript. This research was supported in part by a grant from the North Pacific Research Board (no. R0605). This is publication no. 248 of the North Pacific Research Board. Literature cited Alderdice, D. F., and C. R. Forrester. 1971. Effects of salinity, temperature, and dissolved oxygen on early development of the Pacific cod ( Gadus macrocephalus) . J. Fish. Res. Board Can. 28:883-902. Anderson, P. J., and J. F. Piatt. 1999. Community reorganization in the Gulf of Alaska following ocean climate regime shift. Mar. Ecol. Prog. Ser. 189:117-123. Benoit, H P, and P. Pepin. 1999. Interaction of rearing temperature and maternal influence on egg development rates and larval size at hatch in yellowtail flounder ( Pleuronectes ferrugineus ). Can. J. Fish. Aquat. Sci. 56:785-794. Buckley, L. J., E. M. Caldarone, R. G. Lough, and J. M. St. Onge- Burns. 2006. Ontogenetic and seasonal trends in recent growth ratesof Atlantic cod and haddock larvae on Georges Bank: effects of photoperiod and temperature. Can. J. Fish. Aquat. Sci. 325: 205—226. Christiansen, M. R., and B. Korsgaard. 1999. Protein metabolism, growth and pigmentation patterns during metamorphosis of plaice ( Pleuronectes platessa) larvae. J. Exp. Mar. Biol. Ecol. 237:225- 241. Clemmesen, C., V., Buhler, G. Carvalho, R. Case, G. Evans, L. Hauser, W. F. Hutchinson, O. S. Kjesbu, H. Mempel, E. Moksness, H. Otteraa, H. Paulsen, A. Thorsen, and T. Svaasand. 2003. Variability in condition and growth of Atlantic cod larvae and juveniles reared in mesocosms: environmental and maternal effects. J. Fish Biol. 62:706-723. Colton, A. R., and T. P. Hurst. 2010. Behavioral responses to light gradients, olfactory cues, and prey in larvae of two North Pacific gadids (Gadus macrocephalus and Theragra chalcogramma). Environ. Biol. Fishes 88:39-49. Conceicao, L. E. C., M. Yufera, P. Makridis, S. Morais, and M. T. Dinis. 2010. Live feeds for early stages of fish rearing. Aqua- cult. Res. 41:613-640. Cowen, J. H., and R. F. Shaw. 2002. Recruitment. In Fishery science: the unique contributions of early life stages (L. A. Fuiman and R. G. Werner, eds.), p. 88-111. Blackwell Science Ltd., Oxford. Davis, M. W., and M. L. Ottmar. 2009. Vertical distribution of juvenile Pacific cod Gadus macrocephalus : potential role of light, temperature, food, and age. Aquat. Biol. 8:29-37. Elliott, J. M., and M. A. Hurley. 1995. The functional relationship between body size and growth rate in fish. Funct. Ecol. 9:625-627. Folkvord, A. 2005. Comparison of size-at-age of larval Atlantic cod ( Gadus morhua) from different populations based on size- and temperature-dependent growth models. Can. J. Fish. Aquat. Sci. 62:1037-1052. Folkvord, A., and H. Ottera. 1993. Effects of initial size distribution, day length and feeding frequency on growth, survival and cannibalism Hurst et al.: Growth rates of Gadus macrocephalus 391 in juvenile Atlantic cod (Gadus morhua L.). Aquacul- ture 114:243-260. Freitas, V., J. Campos, M. Fonds, and H. W. van der Veer. 2007. Potential impact of temperature change on epi- benthic predator-bivalve prey interactions in temperate estuaries. J. Therm. Biol. 32:328-340. Green, B. S., and M. I. McCormick. 2005. Maternal and paternal effects determine size, growth and performance in larvae of a tropical reef fish. Mar. Ecol. Prog. Ser. 289:263-272. Hollowed, A. B., N. A. Bond, T. K. Wilderbuer, W. T. Stock- hausen, Z. T. A'Mar, R. J. Beamish, J. E. Overland and M. J. Schirripa. 2009. A framework for modelling fish and shellfish responses to future climate change. ICES J. Mar. Sci. 66:1584-1594. Hollowed, A. B., S. R. Hare, and W. S. Wooster. 2001. Pacific Basin climate variability and patterns of Northeast Pacific marine fish production. Prog. Ocean- ogr. 49:257-282. Houde, E. D. 1996. Evaluating stage-specific survival during the early life of fishes. In Survival strategies in early life stages of marine resources (Y. Watanabe, Y. Yamashita, and Y. Oozcki, eds.), p. 51-65. A. A. Balkema, Rotterdam. Hunt, R. L., and P. J. Stabeno. 2002. Climate change and the control of energy flow in the southeastern Bering Sea. Prog. Oceanogr. 55:5-22. Hurst, T. P., and A. A. Abookire. 2006. Temporal and spatial variation in potential and realized growth rates of age-0 northern rock sole. J. Fish Biol. 68:905-919. Hurst, T. P., A. A. Abookire, and B. Knoth. 2010. Quantifying thermal effects on contemporary growth variability to predict responses to climate change in northern rock sole (Lepidopsetta polyxystra). Can. J. Fish. Aquat. Sci. 67:97-107. Hurst, T. P., D. W. Cooper, J. S. Scheingross, E. M. Seale, B. J. Laurel, and M. L. Spencer. 2009. Effects of ontogeny, temperature, and light on vertical movements of larval Pacific cod ( Gadus macrocephalus). Fish. Oceanogr. 18:301-311. Jones, C. M. 2002. Age and growth. In Fishery science: the unique contributions of early life stages (L. A. Fuiman and R. G. Werner, eds.), p. 33-63. Blackwell Science Ltd., Oxford. Kamler, E. 1992. Early life history of fish: An energetics approach. Chapman & Hall, London. Karikiri, M., R. Berghahn, and H. W. van der Veer. 1991. Variation in settlement and growth of 0-group plaice (Pleuronectes platessa L.) in the Dutch Wadden Sea as determined by otolith microstructure analysis. Neth. J. Sea Res. 27:345-351. Kristiansen, T., 0. Fiksen, and A. Folkvord. 2007. Modeling feeding, growth, and habitat selection in larval Atlantic cod ( Gadus morhua)'. observations and model predictions in a mesocosm environment. Can. J. Fish. Aquat. Sci. 64:136-151. Lanksbury, J. A., J. T. Duffy-Anderson, K. L. Mier, M. S. Busby and P. J. Stabeno. 2007. Distribution and transport patterns of northern rock sole, Lepidopsetta polyxystra , larvae in the south- eastern Bering Sea. Prog. Oceanogr. 72:39-62. Laurel, B. J., T. P. Hurst, L. A. Copeman and M. W. Davis. 2008. The role of temperature on the growth and survival of early and late hatching Pacific cod larvae ( Gadus macrocephalus). J. Plankton Res. 30:1051-1060. Martell, D. J., J. D. Kieffer, and E. A. Trippel. 2005. Effects of temperature during early life history on embryonic and larval development and growth in haddock. J. Fish Biol. 66:1558—1575. Monk, J., V. Puvanendran, and J. A. Brown. 2008. Does different tank bottom colour affect the growth, survival and foraging behaviour of Atlantic cod ( Gadus morhua) larvae? Aquaculture 277:197-202. NMFS (National Marine Fisheries Service). 2008. Fisheries of the United States. Current fishery statistics no. 2007. Office of Science and Technology, Fishery Statistics Div., Silver Spring, MD. Neuman, M. J., D. A. Witting, and K. W. Able. 2001. Relationships between otolith microstructure, otolith growth, somatic growth and ontogenetic tran- sitions in two cohorts of windowpane. J. Fish Biol. 58:967-984. Otterlei, E., G. Nyhammer, A. Folkvord, and S. O. Stefansson. 1999. Temperature- and size-dependent growth of larval and early juvenile Atlantic cod (Gadus morhua)'. a comparative study of Norwegian coastal cod and north- east Arctic cod. Can. J. Fish. Aquat. Sci. 56:2099- 2111. Park, H. G., V. Puvanendran, A. Kellett, C. C. Parrish, and J. A. Brown. 2006. Effect of enriched rotifers on growth, survival, and composition of larval Atlantic cod (Gadus morhua). ICES J. Mar. Sci. 63:285-295. Perry, A. L., P. J. Low, J. R. Ellis, and J. D. Reynolds. 2005. Climate change and distribution shifts in marine fishes. Science 308:1912-1915. Peterson, W. T., and F. B. Schwing. 2003. A new climate regime in northeast Pacific ecosystems. Geophys. Res. Lett. 30 (17), 1896, doi:10.1029/2003GL0157528. Rakocinski, C. F., M. S. Peterson, B. H. Comyns, G. A. Zapfe, and G. L. Fulling. 2006. Do abiotic factors drive the early growth of juvenile spot (Leiostomus xanthurus )? Fish Res. 82:186-193. Rijnsdorp, A. D., M. A. Peck, G. H. Engelhard, C. Mollmann, and J. K. Pinnegar. 2009. Resolving the effect of climate change on fish populations. ICES J. Mar. Sci. 66:1570-1583. Royer, T. C., and C. E. Grosch. 2006. Ocean warming and freshening in the north- ern Gulf of Alaska. Geophys. Res. Lett. 33:L16605, doi:10.1029/2006GL026767. Sigourney, D. B., B. H. Letcher, M. Obedzinski, and R. A. Cunjak. 2008. Size-independent growth in fishes: patterns, models and metrics. J. Fish Biol. 72:2435-2455. Sogard, S. M., and B. L. Olla. 1994. The potential for intracohort cannibalism in age-0 walleye pollock, Theragra chalcogramma, as determined under laboratory conditions. Environ. Biol. Fishes 39:183-190. Stabeno, P. J., N. A. Bond, and S. A. Salo. 2007. On the recent warming of the southeastern Bering Sea shelf. Deep Sea Res. II (Top. Stud. Oceanogr.) 54:2599-2618. Torres, J. J., R. I. Brightman, J. Donnelly, and J. Harvey. 1996. Energetics of larval red drum, Sciaenops ocellatus. 392 Fishery Bulletin 108(4) Part 1: Oxygen consumption, specific dynamic action, and nitrogen excretion. Fish. Bull. 94:756-765. van der Veer, H. W., and J. I. Witte. 1993. The ‘maximum growth/optimal food condition’ hypothesis: a test for 0-group plaice Pleuronectes pla- tessa in the Dutch Wadden Sea. Mar. Ecol. Prog. Ser. 101:81-90. Wijekoon, M. P., V. Puvanendran, D. W. Ings, and J. A. Brown. 2009. Possible countergradient variation in growth of juvenile cod Gadus morhua from the northwest Atlantic. Mar. Ecol. Prog. Ser. 375:229-238. Wootton, R. J. 1990. Ecology of teleost fishes. Chapman & Hall, London. Yatsu, A., K. Y. Aydin, J. R. King, G. A. McFarlane, S. Chiba, K. Tadokoro, M. Kaeriyama, and Y. Watanabe. 2008. Elucidating dynamic responses of North Pacific fish populations to climatic forcing: Influence of life- history strategy. Prog. Oceanogr. 77:252-268. Yufera, M., and M. J. Darias. 2007. The onset of exogenous feeding in marine fish larvae. Aquaculture 268:53-63. 393 Feeding ecology of juvenile Pacific salmon ( Oncorhynchus spp.) in a northeast Pacific fjord: diet, availability of zooplankton, selectivity for prey, and potential competition for prey resources Email address for contact author: sbollens@vancouver.wsu.edu 1 School of Earth and Environmental Sciences and School of Biological Sciences Washington State University 14204 NE Salmon Creek Avenue Vancouver, Washington 98686 2 School of Arts & Sciences Endicott College 376 Hale Street Beverly, Massachusetts 01915 3 School of Aquatic and Fishery Sciences University of Washington 1122 Boat Street Seattle, Washington 98195 4 School of Oceanography Campus Box 357940, University of Washington Seattle, Washington 98195 Abstract — We investigated the feed- ing ecology of juvenile salmon during the critical early life-history stage of transition from shallow to deep marine waters by sampling two sta- tions (190 m and 60 m deep) in a northeast Pacific fjord (Dabob Bay, WA) between May 1985 and October 1987. Four species of Pacific salmon— Oncorhynchus keta (chum), O. tshawytscha (Chinook), O. gorbusclia (pink), and O. kisutch (coho) — were examined for stomach contents. Diets of these fishes varied temporally, spa- tially, and between species, but were dominated by insects, euphausiids, and decapod larvae. Zooplankton assemblages and dry weights differed between stations, and less so between years. Salmon often demonstrated strongly positive or negative selection for specific prey types: copepods were far more abundant in the zooplank- ton than in the diet, whereas Insecta, Araneae, Cephalapoda, Teleostei, and Ctenophora were more abundant in the diet than in the plankton. Overall diet overlap was highest for Chinook and coho salmon (mean=77.9%) — spe- cies that seldom were found together. Chum and Chinook salmon were found together the most frequently, but diet overlap was lower (38.8%) and zooplankton biomass was not correlated with their gut fullness (% body weight). Thus, despite occasional occurrences of significant diet overlap between salmon species, our results indicate that interspecific competition among juvenile salmon does not occur in Dabob Bay. Manuscript submitted 10 September 2009. Manuscript accepted 16 June 2010. Fish. Bull. 108:393-407 (2010). The views and opinions expressed or implied in this article are those of the author (or authors) and do not necessarily reflect the position of the National Marine Fisheries Service, NOAA. Stephen M. Bollens (contact author)1 Rian vanden Hooff 1 Mari Butler2 Jeffery R. Cordell3 Bruce W. Frost4 Juvenile salmon migrating into coastal waters face a variety of challenges as they adjust to a rapidly changing environment, such as outmigration timing, physiological acclimation, new prey fields, new predators, and com- petition for resources (Pearcy, 1992; Magnusson and Hilborn, 2003). High rates of natural mortality have been attributed to the ocean entry transi- tion (Parker, 1971; Bax, 1983), yet we still have a poor understanding of the processes governing this mortal- ity. Several mechanisms have been hypothesized — disruption of freshwa- ter hydrological conditions, degrada- tion of estuarine nursery conditions, and interannual variability of preda- tor abundance and prey resources in the marine environment (Levings and van Densen, 1990; Willette, 2001; Logerwell et ah, 2003). Comprehensive management of salmon fisheries, including their conservation and recovery, requires detailed understanding of juvenile salmon ecology during this critical transition. Increased understanding of feeding relationships and poten- tial density-dependent effects, such as diet overlap among co-occurring species and resource limitation could allow for a more ecologically based approach for rebuilding threatened salmon stocks. Specifically, species in- teractions and their response to vari- ability in prey resources could be im- portant factors for predicting marine survival and the forecasting of adult returns (Logerwell et al., 2003). Juvenile salmon have been docu- mented feeding in a variety of habi- tats, including freshwater (Keeley and Grant, 2001; Hampton et ah, 2006), estuaries (Healey, 1980; Mur- phy et ah, 1988; Reese et al., 2009), and in the coastal ocean (e.g., Bro- deur and Pearcy, 1990; Daly et al., 2009). Inland marine waters (e.g., bays, straits, sounds, fjords, etc.) have been characterized less fre- quently (Sturdevant et al., 2004; Romanuk and Levings, 2005; Saito et al., 2009). Like estuaries, these inland marine salmon habitats are more geographically and ecologically diverse than offshore habitats, and thus it may be more difficult to gen- eralize about juvenile salmon feeding ecology in these areas. Dabob Bay, a temperate marine fjord in Puget Sound, northwestern Washington, has been the site of nu- merous studies of plankton dynamics 394 Fishery Bulletin 108(4) (e.g., Frost, 1988; Bollens et al., 1992a, 1992b; Frost, 2005 and references therein). Although these stud- ies have provided extensive insight into zooplankton population dynamics and predator-prey interactions, none has specifically reported on the seasonality of zoo- plankton community composition from this area. Four juvenile salmon species reside in Dabob Bay temporarily during outmigration to the Pacific Ocean (Bollens and Frost, 1989): Oncorhynchus keta (chum), O. tshawytscha (Chinook), O. gorbuscha (pink), and O. kisutch (coho). Diets of these four species in the fjord have not been described. Feeding habits of juvenile chum salmon in neritic waters have been described for a nearby location in Hood Canal (Simenstad and Salo, 1980) and feed- ing habits of juvenile salmon in other nursery areas of Puget Sound have also been described (Simenstad et al., 1982; Duffy et al., 2005). Much of our understanding of juvenile salmon feed- ing ecology in northeast Pacific marine waters has been based on detailed analyses of stomach contents, but Figure t Map of Dabob Bay, Washington. Four species of juvenile Pacific salmon — chum ( Oncorhynchus keta), Chinook (O. tshawytscha ), coho (O. kisutch ), and pink (O. gorbuscha) — and zooplankton were sampled at two stations between April 1985 and October 1987 to examine salmon diet, zooplank- ton availability, feeding selectivity, and potential competi- tion for prey. D=location of the deep (190-m) central bay tstation, and S=the location of the shallow (60-m) near- shore station. has been limited by a lack of corresponding analyses of prey fields. Some recent studies have identified the importance of prey selectivity as a factor in assess- ing the trophic ecology of salmon during early marine residence (Landingham et al., 1998; Schabetsberger et al., 2003). Furthermore, the highly variable diets demonstrated within and between studies indicate that temporal (seasonal, interannual, and interdecadal) and spatial scales of variability are important, and therefore pose a significant challenge for the design of field stud- ies. The result has been an incomplete understanding of juvenile salmon response to dynamic zooplankton prey fields, particularly during one of the most critical life-history stages, i.e., the early ocean transition phase (Beamish and Mahnken, 2001). Our objectives were 1) to investigate the diet compo- sition of four salmon species ( Oncorhynchus spp.) col- lected from two stations (nearshore-shallow and central- basin-deep) in Dabob Bay over three years and several seasons; 2) to determine salmon feeding selectivity (i.e., their diet in relation to prey availability); and 3) to explore potential resource competition among these species during their early marine residence. Materials and methods Fish collection and processing Between May 1985 and October 1987 we sampled two stations in Dabob Bay, WA (47°45'-50'N lat., 122°50/W long.): a deep (190 m), central station, and a shallow (60 m), nearshore station 9 km apart (Fig. 1). Fish were sampled at night (just after dusk, i.e., when there was no apparent daylight) with two gear types: a midwater trawl with a mouth area of 81.0 m2 (9. 0x9.0 m), and a surface tow net with a mouth area of 18.3 m2 (3. 0x6.1 m) (see Bollens and Frost, 1989 for details). The midwater trawl was towed obliquely from a 50-m depth to the surface at a mean speed of 150 cm/s. Because of concerns about possible avoidance of this net by fish in the upper few meters of the water column (e.g., due to ship wake and propeller wash), we also deployed the surface tow net in the upper three meters of the water column, towed at a mean speed of 80 cm/s behind (50 m) and between (at a midpoint of 50 m) two different vessels (one vessel 5 m and one 15 m in length). Fish were col- lected at each station during each of four seasons (spring: April-May; early summer: June-July; late summer: August; and autumn: October) in each of three years (1985-87), except for spring of 1985, when no fish were collected (Table 1). Fish collected with the midwater or surface trawls were sorted and counted, and the catch was weighed by species, lengths were measured, and then individuals of five predetermined size (fork length [FL] ) classes (<49 mm, 50-74 mm, Bollens et al.: Feeding ecology of juvenile Oncorhynchus spp. 395 Table 1 Number of juvenile Chinook (Oncorhynchus tshawytscha), chum (O. keta ), coho (O. kisutch), and pink (O. gorbuscha) salmon stomach samples and zooplankton samples collected during 12 visits to each of two stations in Dabob Bay, Washington, 1985-87. Sampling dates are grouped by year within season, number of samples for the deep (D, 190 m) station are indicated on the left and those for the shallow (S, 60 m) station indicated on the right. Diet samples were pooled from midwater trawl and surface trawl collections at each station for each date or dates; duplicate zooplankton samples were collected at each station on each date by vertical hauls of plankton nets. Fish samples were not collected in spring of 1985. Number of juvenile salmon stomachs Number of zooplankton samples Chinook Chum Coho Pink Season and sampling dates D S D S D S D S D S Spring 1985 30 Apr and 29 May 1986 0 0 16 15 0 0 3 7 2 2 2 2 6-7 May 1987 0 0 6 10 0 0 0 0 2 2 Subtotal 0 0 22 25 0 0 3 7 6 6 Early summer 19-20 Jun 1985, 24-25 Jun 1985, 26 Jul 1985 5 10 14 9 0 0 0 0 2 2 4-5 Jun 1986 4 3 10 4 5 1 0 0 2 2 17-18 Jun 1987, 13 Jul 1987 3 5 0 14 1 2 0 0 2 2 Subtotal 12 18 24 27 6 3 0 0 6 6 Late summer 19-20 Aug 1985, 26-27 Aug 1985 4 5 10 12 0 0 0 0 2 2 12 Aug 1986, 14 Aug 1986 5 7 7 3 0 0 0 0 2 2 19-20 Aug 1987 11 6 7 0 1 0 0 0 2 2 Subtotal 20 18 24 15 1 0 0 0 6 6 Autumn 7-9 Oct 1985 0 6 5 0 0 0 0 0 2 2 22 Oct 1986 6 10 7 0 0 0 0 0 2 2 14-15 Oct 1987 29 3 3 0 0 0 0 0 2 2 Subtotal 35 19 15 0 0 0 0 0 6 6 Total 67 55 85 67 7 3 3 7 24 24 75-99 mm, 100-149 mm, >150 mm) were removed and 0. 5-3.0 cc of undiluted formaldehyde were injected into their gut cavity to halt digestion. Each fish was then stored in dilute (5%) formalin-seawater solution. Fish from both trawl types were combined to yield a repre- sentative sample of fish residing in the upper 50 m of the water column at night. Subsequently, each fish was weighed (g, wet weight) and measured in the labora- tory, and its stomach excised. Stomach contents were weighed (wet weight), digestion stage was recorded, and then prey were identified to the lowest possible taxon (often to species), and enumerated and weighed (wet weight) by category. Diet composition was summarized as the normalized average percent prey biomass ( [bio- mass of taxonomic group]/[total weight of stomach - the weight of the unidentified portion of the stomach con- tents]). A total of 294 salmon stomachs were examined; six were empty and were not considered further. For ease of comparison, we pooled the original 122 prey taxa observed into 13 categories that made up at least 10% of any one fish’s total normalized prey bio- mass. Data were stratified by salmon species and size class, season, and station across the three years pooled. In most cases the pooled data were a good representa- tion of the three individual years of data, but in some cases there were interesting interannual differences, as discussed below. Collection and processing of zooplankton Zooplankton collections were made at the two stations within 1-2 hours of the fish trawls by using a 1-m2 mouth opening and 333-pm mesh multiple net sampler (Frost, 1988) in 1985 and 1986 and a 1-m diameter mouth opening and 216 pm mesh Puget Sound net (Research Nets Inc., Seattle) (Miller et al., 1977) in 1987. Duplicate zooplankton samples were collected at each station and on each date (Table 1). A subsample (collected with a Stempel pipette) of 1-2% of the ani- mals was taken and the species were enumerated and 396 Fishery Bulletin 108(4) identified for all dates and stations at which fish were collected. Identification of gelatinous animals was gen- erally poor because of their damage during collection. Abundances (densities) were calculated from the sub- sample counts and volumes (m3) of water filtered (as measured by a flow meter). Data from different depth strata were weighted by the size (height, m) of the depth strata and then averages for the upper 50 m were computed. Zooplankton community composition was described by station by using the data from duplicate zooplankton tows to compute average abundances for each taxonomic group for each of four seasons (spring, early summer, late summer, and autumn) over the three years studied (1985-87). Dry weights (g/m2) for zooplankton >216 pm were measured as described in Bollens et al. (1992b). For ease and clarity of presentation of the zooplank- ton compositional data, the category “other Copepoda” that is used in presenting juvenile salmon gut contents was partitioned into several subcategories, i.e., the gen- era Oithona, Metridia, and Pseudocalanus. In addition, several taxonomic groups that were common in salmon gut contents were rarely or never observed in the zoo- plankton samples (e.g., Insecta and Araneae, Cepha- lopoda, Teleostei, and Ctenophora) and were therefore placed into the “other” category for characterizing zoo- plankton community composition. Data analysis We estimated interspecific diet overlap between co-occur- ring juvenile salmon species on the basis of biomass of prey in common, using Schoener’s (1970) percent simi- larity index (PSI): PS/,,y = 10°(l-0.5(^|Pxv-PVii|)), (1) where Pxi = percent biomass of food category ( i ) in the stomach of species x; and Pyi = percent biomass of food category ( i ) in the stomach of species y. We first pooled the prey biomass data by size class for each salmon species. PSI calculations were made for the subset of samples with a minimum of three stomachs per species. We examined intraspecific spa- tial variation in diet by calculating the PSI of juvenile salmon between the two stations. PSI values >60% were considered significant (Brodeur and Pearcy, 1992; Landingham et al., 1998). Prey selectivity by juvenile salmon species was exam- ined by using Ivlev’s (1961) electivity index (Et): E, = {ri-pi)Hri +Pl), (2) where El = the electivity index; r- = the numerical proportion of the ith taxon in the stomachs; and pt = the proportion of the same taxon in the environment. The electivity values provide a species-specific mea- sure of prey selection by allowing a comparison of salm- on gut contents to available prey. Values for Et range from -1 to 1, where 1 indicates the highest selectivity (i.e., present in the diet, but never in the zooplankton samples), and -1 indicates lowest selectivity (i.e., never in the diet, but present in the zooplankton samples). We summarized these observations as average species- specific electivity scores for all size classes, seasons, and years combined to compare selection across salmon species. We compared juvenile salmon gut fullness (% body weight) and zooplankton abundance (dry weight) at each station, as well as each salmon species’ gut fullness be- tween the two stations, using Spearman’s rank correla- tions (Zar, 1999). The difference in mean zooplankton abundance (dry weight) between the two stations was tested by using a Mann-Whitney U test (Zar, 1999). Results Juvenile salmon diet Juvenile salmon in Dabob Bay showed species-specific patterns of occurrence throughout the April-October time period across years. In general, more and smaller juvenile salmon were caught at the nearshore shallow station during spring and early summer, and more and larger juvenile salmon were caught at the central-bay deep station during late summer and autumn. Chum salmon were most prominent during the spring and early summer, whereas Chinook salmon were caught more frequently later in the year, particularly at the deep sta- tion (190 m). All four salmon species were predominantly planktivorous, although there were some exceptions as described below. Because of the greater occurrence of chum and Chinook salmon diet samples, our analyses were focused on these two species. Juvenile chum salmon Chum salmon exhibited striking ontogenetic and spatial variation in diets, including a tendency for fish <100 mm FL to consume more insects and larvaceans and those >100 mm to consume more amphipods and decapod larvae. At the deep station in late spring, euphausiids were the dominant prey type by weight (Fig. 2). Small (<49 mm) fish also consumed insects, arachnids, and copepods other than Calanus pacificus, whereas larger fish (75-99 mm) consumed high percentages of teleosts (primarily unidentified species), hyperiid amphipods, and decapod larvae. The single fish of the 50-74 mm size range captured during early summer at the deep station consumed mostly insects and arachnids, whereas fish between 75 and 150 mm consumed predominantly larvaceans and hyperiids. The single large fish (>150 mm) contained exclusively euphausiids. In late summer at the deep station, fish 100-149 mm consumed mostly hyperiid amphipods, whereas fish >150 mm consumed Bollens et al.: Feeding ecology of juvenile Oncorhynchus spp. 397 T3 OO M LO ~ oo 03 3 d TO co 5 03 O n> b£ ^ d « ° rn Td CD II n a> ^ Jh cd co 73 £ ^ fM & i= g £ cd o cd O cd q; O C co o cd CO CD cd co O' ^ CO r-SZj £ c CO 398 Fishery Bulletin 108(4) Early summer Late summer (/) in ro E o !o c o <5 O Q. E o u aJ Q Autumn A Deep station (190 m) B Shallow station (60 m) ~Y\ Calanus pacificus Other Copepoda [] Insecta and Araneae Y /\ Cephalopoda Ic'/cl Ctenophora Decapoda Euphausiacea | Gammaridea [§§§§] Hyperiidea Larvacea Pteropoda Teleostei Other Size class (mm) Figure 3 Stomach contents of juvenile Chinook salmon (Oncorhynchus tshawytscha) caught at the deep station (A) and shallow sta- tion (B) of Dabob Bay, Washington, during different seasons. Collections were made with a midwater trawl and a surface tow net at each station on each date. Data were pooled over three years (1985-87) for each season and calculated as the normalized biomass of each taxonomic group. mostly euphausiids, and the remainder consumed mainly gammarid and hyperiid amphipods. In autumn, the diet of fish >150 mm from the deep station was com- posed more evenly of several prey categories, including euphausiids, copepods other than C. pacificus, gammarid and hyperiid amphipods, larvaceans, and cephalopods. In contrast, prey of chum salmon <150 mm at the shallow station were more varied, and few euphausiids were present (Fig. 2B). Teleosts and larvaceans, how- ever, were represented in similar proportions to those at the deep station. Larger chum salmon (>100 mm) sampled in early summer were the only group that consumed relatively large amounts of ctenophores; the remainder of their prey consisted mainly of decapod larvae and hyperiids. During late summer at the shal- low station, chum salmon consumed decapod larvae, gammarid amphipods, copepods other than C. pacificus , and larvaceans. Large fish (>150 mm) collected in au- tumn fed almost entirely on euphausiids. Juvenile Chinook salmon There was substantial variation in Chinook salmon diets across size classes, seasons, and stations. At the deep station during early summer, euphausiids consti- tuted a large proportion of the diet of Chinook salmon between 75 and 99 mm (Fig 3A). Fish 100-149 mm consumed a more evenly distributed mix of prey, domi- nated by teleosts and euphausiids. In late summer, fish 100-149 mm at the deep station consumed mostly hyperiid amphipods and euphausiids. In autumn, fish >100 mm at the deep station consumed mostly gam- marids, euphausiids, insects, arachnids, cephalopods, and hyperiids. Bollens et al.: Feeding ecology of juvenile Oncorhynchus spp. 399 At the shallow station in early summer, decapod larvae dominated the diet of Chinook salmon >75 mm (Fig. 3B). During both late summer and autumn, in- sects and arachnids dominated the diets of all three size classes of Chinook salmon, but in autumn, those Chinook salmon >150 mm also consumed hyperiid amphipods. Diet of other salmon species Fewer diet samples were available for juvenile pink and coho salmon than for juvenile chum and Chinook salmon. Diet samples of pink salmon were available only for small fish (<49 mm) in spring. At the deep station these fish contained about 50% euphausiids, and gammarids, copepods other than C. pacificus, insects, and arachnids made up the other 50%. At the shallow station, their diet included a variety of prey consisting mainly of teleosts, pteropods, copepods other than C. pacificus, and fewer insects, decapod larvae, gammarids, and “others.” The “others” in this case were mostly bivalves, whereas the “others” for the co-occurring juvenile chum salmon were mostly chaetognaths. A few juvenile coho salmon were also caught at the two sample stations. At the deep station, diet of coho salmon 100-149 mm consisted primarily of decapod larvae and euphausiids, and a single larger (>150 mm) coho salmon consumed mostly decapod larvae in early summer and another large coho, only gammarid am- phipods in late summer. At the shallow station, three coho salmon were caught in June: one (75-99 mm) con- sumed about 75% “other” taxa (mostly the ostracod Euphilomedes), and the two other fish consumed almost exclusively decapod larvae (like the five co-occurring Chinook salmon). Juvenile salmon gut fullness in relation to zooplankton abundance Juvenile salmon gut fullness (% body weight) was not related to zooplankton abundance (dry weight) (Fig. 4; Spearman’s rank correlation, P>0.05). For all four species of salmon, gut fullness was generally greater in spring and early summer and declined somewhat during late summer and autumn (chum salmon at the deep station in 1985 was an exception to this). In con- trast, zooplankton dry weight at the deep station peaked in the fall, and minima occurred in the spring. At the shallow station, peaks in zooplankton dry weight tended to occur in the summer, and less pronounced minima occurred in spring and autumn. Zooplankton dry weight was substantially greater at the deep station than at the shallow station (P<0.001, Mann-Whitney U test; Zar, 1999), generally by one order of magnitude. Some of this difference may have been due to the three-fold greater water column depth at the deep station, whereas a generally larger part of this difference was due to greater abundances of large-body zooplankton in the deep station samples. In addition, no significant cor- relation was found for individual salmon species’ gut 1985 1986 1987 • Chum O Chinook ▼ Pink a Coho <=) Zooplankton dry weight Figure 4 Zooplankton (>216 pm) dry weights (g/m2) and fish stomach fullness (mean % body weight) for four spe- cies of Pacific salmon — chum ( Oncorhynchus keta), Chi- nook (O. tshawytscha), pink (O. gorbuscha), and coho (O. kisutch) — sampled between April 1985 and October 1987 at a deep station (A) and a shallow station (B) in Dabob Bay, Washington. Fish collections were made with a midwater trawl and a surface tow net, and zoo- plankton collections were made with vertical hauls of a plankton net at each station on each date. Dashed vertical lines separate years. Note the different scales for zooplankton dry weights on the two y axes. fullness at the two different stations (Spearman’s rank correlation, P>0.05; Zar, 1999). Community composition of zooplankton Zooplankton communities (upper 50 m at night) were numerically dominated by copepods at both stations (Fig. 5). The most striking contrast between our zooplankton composition and our juvenile salmonid diet composi- tion was the far greater abundance of copepods in the zooplankton. Substantially different seasonal patterns in zooplankton community composition were observed between stations during 1985 and 1986, but not in 1987. At the deep station, the greatest species richness was typically observed during early spring, whereas two spe- cies— Metridia lucens and Calanus pacficus — dominated 400 Fishery Bulletin 108(4) A Deep station (190 m) 100 80 60 - 40 - 20 - 0 ZZ to 100 C O B Shallow station 80 60 -I 40 20 1 0 £ 1 (60 m) 7 — - ZZ •S 1 /// z* / / ,0.75) and rou- tinely selected against Calanus pacificus, copepods other than C. pacificus, ctenophores, larvaceans and pteropods (£,•<-0.25). All salmon species except pink salmon typi- cally selected for decapod prey (0.25< £,<0.75; both cari- dean shrimp and brachyuran crab larvae were present in the spring, but brachyurans dominated the decapod prey in summer). Hyperiid amphipods and teleosts were generally selected for by chum and Chinook salmon (0.25<£;<0.75) but were consumed in proportion to their relative abundance or were selected against by coho and pink salmon (-0.75<£(<0.25). Euphausiids were gener- ally neutrally selected (-0.25<£;<0.25). However, elec- tivity scores varied substantially within a given salmon species. For example, in 26 Chinook salmon diet samples, electivity indices for euphausiid prey ranged from -1.0 to 0.99 (data not shown), but were not consistently related to predator size, season, or prey abundance. Important ontogenetic changes in prey-selection behavior were revealed by size-specific electivity val- ues (Fig. 6). For instance, smaller size chum salmon strongly avoided larvacean prey, whereas larger chum (>75 mm) showed roughly neutral or positive selectivity. Similarly, small (75-99 mm) Chinook salmon tended to avoid gammarid amphipods, whereas larger individuals (>100 mm) tended to select gammarids. Also, smaller Chinook salmon individuals tended to consume Calanus pacificus at rates proportional to their abundance in the environment, but the larger Chinook salmon strongly avoided these prey. In general, it seemed that larger fish Bollens et al.: Feeding ecology of juvenile Oncorhynchus spp. 401 E E CD A E E h- I O LO Aha!JO0|0 O!|p0ds 0Zj$ E E CD CD I in h- E E CD o o E E o in a" oo ^ ^ LO ^ CO ^ O CO) 03 4_) pa m ~ _D ^ ^ o <2 b ■s aS Q « ^ -d t; ° a ^ CO os O 3 - £ 03 G< G c< 0> G p“3 -7-3 (fl w w -N g -*j 7 O ^ o ^ o> h: a > ■2 ^ ^ ~se G • ’> 03 o * p d) >-< o "C A o3 03 2 £ a* C £ .2 2 G G o c/o -£ _C o +3 OJ o S r-C *© O 3 ^ 5 G G 03 03 rH< o o S W Tr 03 CG 3 r^H N bo w • rH £ 03 CO .5 100 mm in length captured in highly variable near- shore coastal environments have been examined in more recent studies (Moulton, 1997; Landingham et al., 1998; De Robertis et al., 2005, Armstrong et al., 2008). Ours is one of few studies where larger juvenile fish from transitional inland marine habitats have been examined (Willette, 2001; Sturdevant et al., 2004; Armstrong et al., 2008). A variety of zooplanktivorous fish select prey dis- proportionate to their abundance in the environment (e.g., Lazzaro, 1987; Gerking, 1994). For salmon and other fishes, these patterns have been attributed to multiple factors, including: prey size (Brodeur, 1991); prey pigmentation or other visual indicators (Peterson et al., 1982; Schabetsberger et al., 2003); and verti- cal migration behavior of predator and prey (Bollens and Frost, 1989; Viitasalo et al., 2001). Our three-year study of juvenile salmon feeding in Dabob Bay provides additional evidence that juvenile chum, Chinook, coho, and pink salmon exhibit ontogenetic shifts in prey size selection and that they select for larger and more visu- ally conspicuous prey. Previous studies showed that diel vertical migration is an important mediator of plank- tivore trophic interactions in Dabob Bay (e.g., Bollens and Frost, 1989; Frost and Bollens, 1992; Bollens et al., 1993). Juvenile salmon in Dabob Bay used a diverse prey field and demonstrated species-specific prey preferences. Chum and Chinook salmon both highly preferred in- sects, cephalopods, decapod larvae, hyperiid amhipods, Bollens et al.: Feeding ecology of |uvenile Oncorhynchus spp. 403 and teleost prey. Coho and pink salmon both strongly selected for insects, whereas decapod larvae were im- portant to the former and gammarid amphipods to the latter. Our study supports the importance of insect prey to young juvenile salmon in transitional environments (Moulton, 1997; Romanuk and Levings, 2005; Weitkamp and Sturdevant, 2008). The diet of juvenile chum salmon further differed from the other salmon species by the abundance of lar- vaceans (primarily Oikopleura sp.) in their gut contents, and their consumption of ctenophores. These observa- tions are consistent with other reports (Simenstad and Salo, 1980; Black and Low, 1983; Landingham et ah, 1998) and may be related to anatomical gut specializa- tion, which enables chum salmon to assimilate prey items that other salmon cannot digest (Welch, 1997; Arai et al., 2003). Calanoid copepods have been described as a major diet item for juvenile salmon generally (Pearcy, 1992), and for chum and pink salmon specifically (Godin, 1981; Sturdevant et al., 2004). However, despite a diverse and abundant assemblage of copepod species in Dabob Bay, copepods only represented a modest component of our salmon diets and were particularly limited to smaller predators. Instead, juvenile salmon in Dabob Bay were found more often feeding on numerically less abundant macrozooplankton such as euphausiids, hyperiid am- phipods, and decapod larvae. Similarly, Peterson et al. (1982) showed that juvenile coho and Chinook salmon off Oregon fed more on hyperiid amphipods than on the numerically dominant copepods. These results sup- port the results from other studies that indicate that salmon are more likely to feed on larger, more visible prey items (Healey, 1980; Schabetsberger et al., 2003), in which case abiotic factors (e.g., light intensity and turbidity) and biotic processes (e.g., vertical migration and predator evasion behavior) will be important vari- ables that will help determine stomach fullness and feeding success. Ontogenetic diet thresholds for juvenile salmon at approximately 80 mm, before which teleost prey are less important, have been indicated by other studies (Brodeur, 1991; Keeley and Grant, 2001). In contrast, our electivity results provide evidence that teleost prey were strongly selected for by small (<75 mm) chum and pink salmon during spring, when fish larvae may have been particularly small (Bollens et al., 1992a; Fulmer and Bollens 2005). Simenstad and Salo (1980) found that juvenile chum salmon transitioned from nearshore habitats with epibenthic food sources to neritic habi- tats with pelagic and nektonic food sources when they reached approximately 45-55 mm FL. In other studies, seasonal variability in salmon gut contents has been at- tributed to ontogenetic shifts in feeding preferences or feeding behavior (Beacham, 1986; Brodeur, 1991; Daly et al., 2009). Similarly, our results indicate that small chum, Chinook, and coho salmon select small prey, then larger prey as the fish develop. Small Chinook and coho salmon selected Calanus pacificus roughly in proportion to its abundance, but other copepods were avoided. Similarly, only small coho salmon selected gammarid and hyperiid amphipods. In contrast, only large Chinook salmon selected for gammarids. Thus, both species-specific and ontogenetic shifts in prey pref- erence were observed. Our diet and electivity results should be interpret- ed cautiously because our samples were pooled across broad size, temporal, and spatial scales, and because of limitations associated with sample sizes, net sam- pling biases, and pooling of prey species and life history stages. For example, the range of euphausiid electivity values observed may be due to the pooling of euphausiid species and life-history stages, potentially obscuring euphausiid prey selection patterns observed in other studies (Schabetsberger et al., 2003). Another major caution concerns our ability to deter- mine “available prey” with plankton nets. More mo- bile and larger nektonic prey, such as cephalopods and young fish, are able to avoid conventional plankton nets, with the consequence that electivity indices for these prey types would be biased upward. Conversely, small prey types that are unable to avoid the plankton net (e.g., small copepods) would be proportionately over- represented in the net samples, with the consequence that electivity indices for these forms would be biased downward. We recommend that further research be undertaken into adequately sampling macrozooplank- ton and micronekton (e.g., Gewant and Bollens, 2005), such that a broader and potentially more appropriate range of potential prey for fishes can be quantitatively sampled. Several additional complicating factors should be considered when interpreting electivity indices. First, strongly positive electivity (e.g., E; = 1.00) often results from a rare presence of a species in the gut contents and a corresponding absence of that same species in the plankton. In some cases zero abundance in the plankton may be due to low-volume plankton hauls which under-sample the available prey field. Conversely, an Ex of —1.00 could result from a rare (but nonzero) occurrence of a species in the environment, combined with its absence from the gut contents, perhaps simply because of a low probability of encounter between preda- tor and prey. A final caution concerns the vertical resolution of sampling. Landingham et al. (1998) used both neuston and oblique plankton tows and showed that salmon diet most closely resembled that of the neuston assem- blage. The upper 50 m were sampled with our sampling methods and therefore electivity values may have been biased. For example, if juvenile salmon are primarily feeding near the surface, abundant zooplankton (i.e., co- pepods) that are more deeply distributed may not fully be part of the “available” prey community. We recom- mend finer-scale, vertically resolved sampling of juve- nile salmon and their potential prey in future studies. Dabob Bay has been the site of numerous studies for which the interactions between planktivorous fishes and 404 Fishery Bulletin 108(4) the behaviors exhibited by their potential prey have been explored (Ohman, 1986; Frost, 1988; Bollens and Frost, 1989; Bollens et al., 1992a; Bollens et ah, 1993). Field studies by Bollens and Frost (1989) indicated that abundances of actively feeding planktivorous fish (including Oncorhynchus spp.) are directly linked with the strength and timing of vertical migration exhibited by the copepod Calanus pacificus. Our results indicate that the adaptive response exhibited by species such as Calanus pacificus seems to be an effective mecha- nism for avoiding predation by species such as juvenile salmon. Thus, “available” prey items are not only those that are abundant or of the desired size, but those that are also available for visual detection. Availability may be affected by the prey’s presence or absence from the photic zone, or by the presence of pigmentation that makes the prey more detectable visually. Most of the prey items that were consistently consumed by salmon in this study (e.g., euphausiids, hyperiid and gammarid amphipods, and decapod larvae) possess characteristi- cally dark or large eyes. The ability of salmon to detect these potential prey items may be increased by heavy pigmentation, large body size, and their frequently not- ed association with the near surface layer (Lough, 1976; Peterson et al., 1982) where salmon typically feed. Diet overlap and potential interspecific competition among salmon species A variety of studies have relied on diet overlap as a pri- mary indicator of potential resource competition between co-occurring species. Although there is little consensus among studies of salmon, diet overlap has been most frequently observed between Chinook and coho salmon in Oregon and Washington (Peterson et al., 1982; Emmett et al., 1986; Brodeur and Pearcy, 1990; Brodeur, 1991), and to a lesser extent between chum, pink, and sockeye salmon ( Oncorhynchus nerka) in British Columbia and Southeast Alaska (Healey, 1980; Beacham, 1993; Land- ingham et al., 1998). Our results from Dabob Bay show the greatest spatial and temporal overlap between chum and Chinook salmon but also provide evidence only for resource partitioning (low diet overlap) between these two species, not necessarily competition (which would require resource limitation). In contrast, our data show significant diet overlap between Chinook and coho salmon (average PSI=77.9%), supporting earlier reports of potential resource com- petition between these two species. Although Brodeur and Pearcy (1990) did not see evidence for significant overlap between four salmon species when all observa- tions were combined, they observed significant overlap between Chinook and coho salmon during May and June, as well as during the 1983 El Nino. Likewise we found that significant overlap between Chinook and coho salmon occurred during June (of 1986 and 1987), largely because of the shared consumption of decapod larvae, which are visually conspicuous and seasonally abundant at this time of year. However, the co-occur- rence of juvenile Chinook and coho in Dabob Bay was less prominent than in coastal Oregon (Peterson et al., 1982; Brodeur and Pearcy, 1990; Brodeur and Pearey, 1992); however, our results are based on far fewer data. Despite the co-occurrence of different juvenile salmon species in Dabob Bay, and the occasional occurrence of significant diet overlap between these species, we did not see any indication of food limitation. That is, there was never a significant relationship between stomach fullness and zooplankton biomass, as might be expected if food was limited. However, just as with the electivity indices discussed above, we caution that our vertical plankton net hauls may not adequately sample the po- tential prey of juvenile salmon. Testing for food limita- tion by correlating salmon stomach fullness and the abundance of potentially more appropriate prey (e.g., macrozooplanktonic, micronektonic, and neustonic prey) would prove interesting, but was not possible given our sampling method. Similarly, our comparison of zoo- plankton dry weights with salmon stomach wet weights complicates the interpretation of food limitation and po- tential competition because conversions from wet-weight to dry-weight would be expected to vary between prey taxa (e.g., between gelatinous and crustacean prey). Resource limitation by juvenile salmon during their early marine transition may be influenced by several other factors not addressed in our study, including di- rect and indirect effects of hatchery production in the region (Quinn et al., 2005) and potential diet overlap with other zooplanktivores (Purcell and Sturdevant, 2001). Furthermore, zooplankton dynamics in temper- ate marine waters are clearly influenced by interan- nual (El Nino cycles) and interdecadal (Pacific Decadal Oscillation) scales of climate variability (Mackas et al., 2001; Hooff and Peterson, 2006) and there are im- portant linkages to salmon survival during multiple life-history stages (Beamish and Bouillon, 1993; Loger- well et al., 2003). Although diet data from our three- year study could be averaged across years (as opposed to size classes, etc.), interannual climate factors can- not be overlooked. Indeed, based on a multivariate El Nino-Southern Oscillation index (MEI), 1987 ranks as a moderate to strong El Nino year (April-October MEI average = 1.91), and likewise represents the most anomalous year of our study for seasonal plankton com- position in Dabob Bay. This was particularly apparent at the deeper station, where the abundance of Oithona sp. and larvaceans seemed to be more characteristic of the shallow, nearshore station in 1985 and 1986. Mechanisms underlying the interannual variability of zooplankton composition in Dabob Bay warrant further exploration. Our combination of detailed analyses of prey fields and fish diets is clearly only one approach to under- standing juvenile salmon feeding ecology. Biochemical methods of studying energetics and feeding relation- ships (e.g., Johnson and Schindler, 2009) provide ad- ditional insight into juvenile salmon trophic dynamics. Studies across the variety of habitats encountered by salmon during their outmigration and early residence in marine environments will be necessary to fully un- Bollens et al.: Feeding ecology of juvenile Oncorhynchus spp. 405 derstand species-specific responses to resource and en- vironmental variability. Our results from Dabob Bay indicate that periodic high diet overlap between salmon species may occur. However, evidence of resource parti- tioning, especially between frequently co-occurring spe- cies (e.g., chum and Chinook salmon), combined with a lack of evidence for food limitation (although this should be more explicitly tested in the future), indicates that competition between juvenile salmon is unlikely to oc- cur in this marine fjord. Acknowledgments We thank D. Thoreson, S. Jonasdottir, C. Mobley, and the crew of the RV Barnes for assistance with field work, W. Peterson for assistance with zooplankton analyses, and J. Breckenridge and J. Emerson for assistance with graphics. R. Brodeur and three anonymous reviewers provided comments that substantially improved the manuscript. Special thanks go to S. Simenstad and K. Fresh for helping a young oceanography graduate student bridge the gap between fisheries and biological oceanography. We also thank the University of Otago’s Departments of Zoology and Marine Sciences for provid- ing office space and other support to S. Bollens during the writing of this article. This research was supported by National Science Foundation grant 84-08929 to B. Frost, and Washington State University institutional funds (including a sabbatical leave) made available to S. Bollens. Literature cited Arai, M. N., D. W. Welch, A. L. Dunsmuir, M. C. Jacobs, and A. R. Ladouceur. 2003. Digestion of pelagic Ctenophora and Cnidaria by fish. Can. J. Fish. Aquat. Sci. 60:825-829. Armstrong, J. L., K. W. Myers, D. A. Beauchamp, N. D. Davis, R. V. Walker, J. L. Boldt, J. J Piccolo, L. J. Haldorson, and J. H. Moss. 2008. Interannual and spatial feeding patterns of hatch- ery and wild juvenile pink salmon in the Gulf of Alaska in years of low and high survival. Trans. Am. Fish Soc. 137:1299-1316. Bax, N. J. 1983. Early marine mortality of marked juvenile chum salmon ( Oncorhynchus keta ) released into Hood Canal, Puget Sound, Washington, in 1980. Can. J. Fish. Aquat. Sci. 40:426-435. Beacham, T. D. 1986. Type, quantity, and size of food of Pacific salmon ( Oncorhynchus spp.) in the strait of Juan de Fuca, Brit- ish Columbia. Fish. Bull. 84:77-89. 1993. Competition between juvenile pink (Oncorhyn- chus gorbuscha) and chum salmon (Oncorhynchus keta) and its effect on growth and survival. Can. J. Zool. 71:1270-1274. Beamish, R. J., and D. R. Bouillon. 1993. Pacific salmon production in relation to climate. Can. J. Fish. Aquat. Sci. 50:1002-1016. Beamish, R. J., and C. Mahnken. 2001. A critical size and period hypothesis to explain natural regulation of salmon abundance and the link- age to climate and climate change. Prog. Oceanogr. 49:423-437. Black, F. A., and C. J. Low. 1983. Ctenophores in salmon diets. Trans. Am. Fish. Soc. 112:728. Bollens, S. M., and B. W. Frost. 1989. Zooplanktivorous fish and variable diel vertical migration in the marine planktonic copepod Calanus pacificus. Limnol. Oceanogr. 34:1072-1083. Bollens, S. M„ B. W. Frost, and T. S. Lin. 1992a. Recruitment, growth, and diel vertical migra- tion of Euphausia pacifica in a temperate fjord. Mar. Biol. 114:219-228. Bollens, S. M., B. W. Frost, H. R. Schwaninger, C. S. Davis, K. L. Way, and M. C. Landsteiner. 1992b. Seasonal plankton cycles in a temperate fjord, and comments on the match-mismatch hypothesis. J. Plankton Res. 14:1279-1305. Bollens, S. M., B. W. Frost, K. Osgood, and S. D. Watts. 1993. Vertical distributions and susceptibilities to ver- tebrate predation of the marine copepods Metridia lucens and Calanus pacificus. Limnol. Oceanogr. 38:1841-1851. Brodeur, R. D. 1991. Ontogenetic variations in the type and size of prey consumed by juvenile coho, Oncorhynchus kisutch , and chinook, O. tshawytscha, salmon. Environ. Biol. Fish. 30:303-305. 1992. Factors related to variability in feeding intensity of juvenile coho salmon and chinook salmon. Trans. Am. Fish. Soc. 121:104-114. Brodeur, R. D., and W. G. Pearcy. 1990. Trophic relations of juvenile Pacific salmon off the Oregon and Washington coast. Fish. Bull. 88:617- 636. 1992. Effects of environmental variability on trophic interactions and food web structure in a pelagic upwell- ing ecosystem. Mar. Ecol. Prog. Ser. 84:101-119. Daly, E. A., R. D. Brodeur, and L. A. Weitkamp. 2009. Ontogenetic shifts in diets of juvenile and subadult coho and Chinook salmon in coastal marine waters: important for marine survival? Trans. Am. Fish. Soc. 138:1420-1438. De Robertis, A., C. A. Morgan, R. A. Schabetsberger, R. W. Zabel, R. D. Brodeur, R. L. Emmett, C. M. Knight, G. K. Krut- zikowsky, and E. Casillas. 2005. Columbia River plume fronts. II. Distribution, abundance, and feeding ecology of juvenile salmon. Mar. Ecol. Prog. Ser. 299:33-44. Duffy, E. F., D. A. Beauchamp, and R. M. Buckley. 2005. Early marine life history of juvenile Pacific salmon in two regions of Puget Sound. Estuar. Coast. Shelf Sci. 64:94-107. Emmett, R. L., D. R. Miller, and T. H. Blahm. 1986. Food of juvenile chinook, Oncorhynchus tsawtscha, and coho, O. kisutch , salmon off the northern Oregon and southern Washington coasts, May-September 1980. Calif. Fish Game 72:38-46. Frost, B. W. 1988. Variability and possible adaptive significance of diel vertical migration in Calanus pacificus, a planktonic marine copepod. Bull. Mar. Sci. 43:675-695. 406 Fishery Bulletin 108(4) 2005. Diatom blooms and copepod responses: Paradigm or paradox? Prog. Oceanogr. 67:283-285. Frost, B. W., and S. M. Bollens. 1992. Variability of diel vertical migration in the marine planktonic copepod Pseudocalanus neiumani in relation to its predators. Can. J. Fish Aquat. Sci. 49:1137-1141. Fulmer, J., and S. M. Bollens. 2005. Responses of the chaetognath, Sagitta elegans, and larval Pacific hake, Merluccius productus, to spring diatom and copepod blooms in a temperate fjord (Dabob Bay, Washington, USA). Prog. Ocean- ogr. 67:442-461. Gerking, S. D. 1994. Feeding variability, feeding ecology of fish, p. 41-53. Academic Press, San Diego, CA. Gewant, D. S., and S. M. Bollens. 2005. Macrozooplankton and micronekton of the lower San Francisco Estuary: Seasonal, interannual and regional variation in relation to environmental conditions. Estuaries 28:473-485. Godin, J. J. 1981. Daily patterns of feeding behavior, daily rations, and diets of juvenile pink salmon ( Oncorhynchus gor- buscha ) in two marine bays of British Columbia. Can. J. Fish. Aquat. Sci. 37:10—15. Hampton, S. E., P. Romare, and D. E. Seiler. 2006. Environmentally controlled Daphnia spring increase with implications for sockeye salmon fry in Lake Wash- ington, USA. J. Plankton Res. 28:399-406. Healey, M. C. 1980. Utilization of the Nanaimo River estuary by juve- nile chinook salmon, Oncorhynchus tshawytscha. Fish. Bull. 77:653-668. Hooff, R. C., and W. T. Peterson. 2006. Copepod biodiversity as an indicator of changes in ocean and climate conditions of the northern Califor- nia Current ecosystem. Limnol. Oceanogr. 51:2607- 2620. Ivlev, V. S. 1961. Experimental ecology of the feeding of fishes, 302 p. Yale Univ. Press, New Haven, CT. Johnson, S. P., and D. E. Schindler. 2009. Trophic ecology of Pacific salmon ( Oncorhyn- chus spp.) in the ocean: a synthesis of stable isotope research. Ecol. Res. 24:855-863. Kaczynski, V. W., R. J. Feller, J. Clayton, and R. J. Gerke. 1973. Trophic analysis of juvenile pink and chum salmon ( Oncorhynchus gorbuscha and O. keta) in Puget Sound. J. Fish. Res. Board Canada 30(7):1003-1008. Keeley, E. R., and J. W. A. Grant. 2001 . Prey size of salmonid fishes in streams, lakes, and oceans. Can. J. Fish. Aquat. Sci. 58:1122-1132. Landingham, J., M. V. Sturdevant, and R. D. Brodeur. 1998. Feeding habits of juvenile Pacific salmon in marine waters of southeastern Alaska and northern British Columbia. Fish. Bull. 96:285—302. Lazzaro, X. 1987. A review of planktivorous fishes: their evolution, feeding behaviours, selectivities, and impacts. Hydro- biologia 146:97-167. Levings, C. D., and W. L. T. van Densen. 1990. Strategies for fish habitat management in estu- aries: comparison of estuarine function and fish sur- vival. In Management of freshwater fisheries (W. L. T. van Densen, B. Steinmetz, and R. H. Hughes, eds.), p. 582-593. Proceedings of a symposium organized by the European Inland Fisheries Advisory Commis- sion, Goteborg, Sweden, 31 May-3 June 1988, Pudoc, Wageningen. Logerwell, E. A., N. Mantua, P. W. Lawson, R. C. Francis, and V. N. Agostini. 2003. Tracking environmental processes in the coastal zone for understanding and predicting Oregon coho ( Oncorhynchus kisutch) marine survival. Fish. Ocean- ogr. 12:554-568. Lough, R. G. 1976. Larval dynamics of the Dungeness crab, Cancer magister, off the central Oregon coast, 1970-71. Fish. Bull. 74:353-376. Mackas, D. L., R. E. Thomson, and M. Galbraith. 2001. Changes in the zooplankton community of the British Columbia continental margin, 1985-1999, and their covariation with oceanographic conditions. Can. J. Fish. Aquat. Sci. 58:685-702. Magnusson, A., and R. Hilborn. 2003. Estuarine influence on survival rates of coho ( Oncorhynchus kisutch) and chinook salmon ( Oncorhyn- chus tshawytscha) released from hatcheries on the U.S. Pacific coast. Estuaries 26:1094-1103. Miller, B. S., C. A. Simenstad, L. L. Moulton, K. L. Fresh, F. C. Funk, W. A. Karp, and S. F. Borton. 1977. Puget Sound baseline program: nearshore fish survey. Final Report FRU-UW-7710, 220 p. Univ. Washington, Fish. Res. Institute, Seattle, WA. Moulton, L. L. 1997. Early marine residence, growth, and feeding by juvenile salmon in Northern Cook Inlet, Alaska. Alaska Fish. Res. Bull. 4:154-177. Murphy, J. M., J. F. Thedinga, and K. V. Koski. 1988. Size and diet of juvenile Pacific salmon during seaward migration through a small estuary in south- eastern Alaska. Fish. Bull. 86:213-218. Ohman, M. D. 1986. Predator-limited population growth of the copepod Pseudocalanus sp. J. Plankton Res. 8:673-713. Parker, R. R. 1971. Size selective predation among juvenile salmonid fishes in a British Columbia Inlet J. Fish. Res. Board Can. 28:1503-1510. Pearcy, W. G. 1992. Ocean ecology of North Pacific salmonids, 170 p. Washington Sea Grant, Seattle, WA. Peterson, W. T., R. D. Brodeur, and W. G. Pearcy. 1982. Food habits of juvenile salmon in the Oregon coastal zone, June 1979. Fish. Bull. 80:841-851. Purcell, J. E., and M. V. Sturdevant. 2001. Prey selection and dietary overlap among zooplank- tiverous jellyfish and juvenile fishes in Prince William Sound, Alaska. Mar. Ecol. Prog. Ser. 210:67-83. Quinn, T. P, B. R. Dickerson, and L. A. Vollestad. 2005. Marine survival and distribution patterns of two Puget Sound hatchery populations of coho ( Oncorhyn- chus kisutch) and chinook (Oncorhynchus tshawytscha) salmon. Fish. Res. 76:209-220. Reese, C., N., Hillgruber, M. Sturdevant, A. Wertheimer, W. Smoker, and R. Focht. 2009. Spatial and temporal distribution and the poten- tial for estuarine interactions between wild and hatch- ery chum salmon (Oncorhynchus keta) in Taku Inlet, Alaska. Fish. Bull. 107:433-450. Romanuk, T. N., and C. D. Levings. 2005. Stable isotope analysis of trophic position and ter- Bollens et al.: Feeding ecology of |uvenile Oncorhynchus spp 407 restrial vs. marine carbon sources for juvenile Pacific salmonids in nearshore marine habitats. Fish. Man- agement Ecol. 12:113-121. Saito, T., I. Shimizu, J. Seki, and K. Nagasawa. 2009. Relationship between zooplankton abundance and the early marine life history of juvenile chum salmon Oncorhynchus keta in eastern Hokkaido, Japan. Fish. Sci. 75:303-316 Schabetsberger, R., C. A. Morgan, R. D. Brodeur, C. L. Potts, W. T. Peterson, and R. L. Emmett. 2003. Prey selectivity and diel feeding chronology of juvenile chinook ( Oticorhynchus tshawytscha) and coho (O. kisutch ) salmon in the Columbia River plume. Fish. Oceanogr. 12:523-540. Schoener, T. W. 1970. Nonsynchronous spatial overlap of lizards in patchy habitats. Ecology 51:408-418. Simenstad, C. A., K. L. Fresh, and E. O. Salo. 1982. The role of Puget Sound and Washington coastal estuaries in the life history of Pacific salmon: an unap- preciated function. In Estuarine comparisons (V. S. Kennedy, ed.), p. 343-364. Academic Press, New York. Simenstad, C. A., and E. O. Salo. 1980. Foraging success as a determinant of estuarine and nearshore carrying capacity of juvenile chum salmon (Oncorhynchus keta) in Hood Canal, Washington. In Proceedings of the North Pacific Aquaculture Sym- posium, Report 82-2, p. 21-37. Alaska Sea Grant, Anchorage, AK. Sturdevant, M. V., E. A. Fergusson, J. A. Orsi, and A. Wertheimer. 2004. Diel feeding and gastric evacuation of juvenile pink and chum salmon in Icy Strait, Southeastern Alaska, May-September 2001. N. Pac. Anadr. Fish Comm. Tech. Rep. 5:107-109. Sturdevant, M. V., A. C. Wertheimer, and J. L. Lum. 1996. Diets of juvenile pink and chum salmon in oiled and non-oiled nearshore habitats in Prince William Sound, 1989 and 1990. Am. Fish. Soc. Symp. 18:578-592. Viitasalo, M., J. Flinkman, and M. Viherluoto. 2001. Zooplanktivory in the Baltic Sea: a comparison of prey selectivity by Clupea harengus and Mysis mixta , with reference to prey escape reactions. Mar. Ecol. Prog. Ser. 216:191-200. Weitkamp, L. A., and M. V. Sturdevant. 2008. Food habits and marine survival of juvenile Chi- nook and Coho salmon from marine waters of Southeast Alaska. Fish. Oceanogr. 17:380-395. Welch, D. W. 1997. Anatomical specialization in the gut of Pacific salmon ( Oncorhynchus ): evidence for oceanic limits to salmon production. Can. J. Zool. 75:936-942. Willette, T. M. 2001. Foraging behaviour of juvenile pink salmon (Oncorhynchus gorbuscha) and size-dependent preda- tion risk. Fish. Oceanogr. 10:110-131. Zar, J. H. 1999. Biostatistical analysis, 4th ed., 929 p. Prentice Hall, Upper Saddle River, NJ. 408 Abstract — Despite its recreational and commercial importance, the move- ment patterns and spawning habitats of winter flounder (Pseudopleuronectes americanus) in the Gulf of Maine are poorly understood. To address these uncertainties, 72 adult winter flounder (27-48 cm) were fitted with acoustic transmitters and tracked by passive telemetry in the southern Gulf of Maine between 2007 and 2009. Two sympatric contingents of adult winter flounder were observed, which exhib- ited divergent spawning migrations. One contingent remained in coastal waters during the spawning season, while a smaller contingent of winter flounder was observed migrating to estuarine habitats. Estuarine resi- dence times were highly variable, and ranged from 2 to 91 days (mean = 28 days). Flounder were nearly absent from the estuary during the fall and winter months and were most abundant in the estuary from late spring to early summer. The observed seasonal movements appeared to be strongly related to water temperature. This is the first study to investigate the seasonal distribution, migra- tion, and spawning behavior of adult winter flounder in the Gulf of Maine by using passive acoustic telemetry. This approach offered valuable insight into the life history of this species in nearshore and estuarine habitats and improved the information available for the conservation and management of this species. Manuscript submitted 4 February 2010. Manuscript accepted 30 June 2010. Fish. Bull. 108:408-419 (2010). The views and opinions expressed or implied in this article are those of the author (or authors) and do not necessarily reflect the position of the National Marine Fisheries Service, NOAA. Movement patterns of winter flounder (Pseudopleuronectes americanus ) in the southern Gulf of Maine: observations with the use of passive acoustic telemetry Gregory R. DeCelles (contact author) Steven X. Cadrin Email address for contact author: gdecelles@umassd.edu National Oceanic and Atmospheric Administration and University of Massachusetts Cooperative Marine Education and Research Program School for Marine Science and Technology 200 Mill Rd„ Suite 325 Fairhaven, Massachusetts 02719 Winter flounder ( Pseudopleuronectes americanus) is a commercially and recreationally important flatfish spe- cies that is distributed along the east coast of North America from North Carolina northward to Newfoundland (Bigelow and Schroeder, 1953). Within U.S. waters, this species is managed as three stock units: the southern New England and Mid-Atlantic (SNE-MA); Gulf of Maine; and Georges Bank. To date, there has been less research on the Gulf of Maine stock than on the SNE-MA stock. Seasonal distribution patterns of winter flounder in the Gulf of Maine have emerged from fisheries catch data (McCracken, 1963), mark recapture experiments (Perlmutter, 1947; Howe and Coates, 1975) and annual trawl surveys. However, these sources of data offer limited insight into patterns of estuarine residence and nearshore habitat use for winter flounder because trawl fisheries and research surveys in the Gulf of Maine are typically limited to coastal and offshore waters. In addition, these tra- ditional sampling sources for obtain- ing distribution data cannot identify individual variability in movement or spawning behavior. In the present study, we sought to investigate spawn- ing patterns and seasonal distribu- tion of winter flounder in nearshore and estuarine waters in the southern Gulf of Maine using passive acoustic telemetry. The magnitude of variability in in- dividual migration patterns remains poorly understood for even the most economically valuable fish species (Able and Grothues, 2007), in part because we have yet to recognize the importance of contingent behavior. Contingents are cohesive groups of fish within a population that ex- hibit a common migration pattern (Cadrin and Secor, 2009). Acoustic telemetry offers new opportunities to investigate the fine-scale movements and behavior of individual fish. This technology allows for the recognition of contingent groups within popula- tions (Secor, 1999) and offers insight into the amount of connectivity be- tween estuarine and coastal popula- tions (Able, 2005). Acoustic telem- etry has been used to gather high resolution data on the behavior of juvenile winter flounder in the Gulf of Maine (Fairchild et al., 2009). However, before the present study, this technology had not been used to study adult winter flounder in the Gulf of Maine. Traditionally, winter flounder spawning in the Gulf of Maine was thought to be restricted to estuarine habitats (Bigelow and Schroeder, 1953). This assumption was subse- quently reflected in designations of Essential Fish Habitat (EFH) for this species (NMFS, 1999). How- ever, other observations indicate that winter flounder in the Gulf of Maine also use coastal waters as spawning grounds (i.e., Howe and Coates, 1975). If a proportion of in- DeCelles and Cadrin: Movement patterns of Pseudopleuronectes americanus in the southern Gulf of Maine 409 70“42'W 70°40'W 70‘38'W 70‘36'W 70'341/V Map of the study site showing the locations of Plymouth Bay and Plymouth Estuary in the Gulf of Maine where the movements of winter flounder (Pseudopleuronectes americanus) were tracked with passive acoustic telemetry. Locations of tag releases and acoustic receivers are shown. Receivers used to track winter flounder movements from November 2007 to November 2008 are shown as solid black circles and receivers added to the array in December 2008 are shown as open circles. Circles are not drawn to scale and do not represent the detection radii of acoustic receivers. dividuals spawn in coastal, rather than estuarine habitats, these fish would represent contingents within the larger population. Contingents may provide populations with enhanced stability and resilience because of the variable survival rates of early life history stag- es in different habitats (Secor, 2007; Kerr et al., 2010). Divergent spawning migrations increase the distribution of eggs and larvae, which may reduce the probability of a failed recruitment event (Lambert, 1990). For winter flounder that spawn demersal eggs, the conditions encountered by the early life history stages are largely determined by the locations where the adults spawn. Therefore, divergent spawning migrations may have a large effect on the recruitment success of local populations. We chose to monitor adult winter flounder in Plymouth Bay and the ad- jacent Plymouth Harbor, Kingston Bay, and Duxbury Bay estuary. Plymouth Harbor, Kingston Bay, and Duxbury Bay estuary are commonly referred to as the Plymouth Estuary, and that name will be used throughout this ar- ticle. Winter flounder is a dominant member of the groundfish community in the Plymouth Estuary, and the estuary serves as an important nursery area for this species (Lawton et ah, 1984). Winter flounder are known to spawn in the Plymouth Estuary, where peak spawning occurs in March and April (Entergy1). In addition, Plymouth Bay has been identified as an area where coastal spawning likely occurs (NMFS, 1999). The goal of this study was to obtain high-resolution data on the seasonal distribution, migrations, and spawning behavior of winter flounder in the southern Gulf of Maine. In particular, three primary objectives were addressed during this study: 1) to determine if adults in this region exhibit divergent spawning migrations; 2) to investigate the seasonal distribution of winter flounder in the region, in association with water tem- perature; and 3) to compare the seasonal distribution observed in this study to that observed in past research in the Gulf of Maine where static sampling methods were used. 1 Entergy. 2001. Ichthyoplankton entrainment monitoring at Pilgrim Nuclear Station, January-December, 2000. In Marine ecology studies related to operation of Pilgrim Station. Report no. 57, 65 p. Entergy Nuclear Generation Company. Plymouth, MA. Materials and methods Study site The Plymouth Estuary and Plymouth Bay are located in the southern portion of the Gulf of Maine (Fig. 1). The Plymouth Estuary is bordered on its seaward side by two barrier beaches. Tidal exchange between the Plymouth Estuary and Plymouth Bay occurs through a 2020-m opening between Saquish Head and the northern extrem- ity of Plymouth Beach. The estuary is well mixed and approximately 66% of the water is replaced during each tidal cycle, creating strong tidal currents (Davis, 1984). Plymouth Bay is bordered by Cape Cod Bay to the east. The average depth within Plymouth Estuary is 3.3 m at mean high water and 2.1 m at mean low water, and 410 Fishery Bulletin 108(4) Figure 2 An example of the external attachment method that was used to secure the acoustic transmitters to winter flounder (Pseudopleuronectes americanus) to monitor their movements in the southern Gulf of Maine. The Vemco V92L transmitters were fitted into a harness of 9/16” soft latex tubing and secured in the harness with two-part epoxy (A). Two nickel tagging pins were passed upwards through the blind side of the fish, through the dorsal musculature (B). salinity in the estuary ranges from 0.0 to 33.0 (Iwano- wicz et al.2). The bathymetry of the estuary is complex, and extensive sand and mud flats are exposed at low tide when the surface area of the estuary is reduced front 10,057 acres to 5465 acres (Iwanowicz et al.2). A deeper channel is present between Saquish Head and Plymouth Beach, where depths reach nearly 26 m. There are four main sources of freshwater input to the Plymouth Estuary: the Back River, Bluefish River, Jones River, and Eel River. Acoustic telemetry Adult flounder were captured by using either a small otter trawl, a commercial trawl vessel, or by hook and line. Sampling was nonrandom, and sampling locations were chosen from areas where adult winter flounder have historically been abundant. Tow times with the small otter trawl varied from 10 to 30 minutes, and tows with the commercial vessel were 30 minutes in duration. The total length of each fish was measured to the near- est centimeter, and only adult fish (>27 cm) in good or excellent condition were fitted with acoustic tags. After tagging, fish were held onboard for observation until they were deemed healthy enough for release. All fish were released in close proximity to the capture site in order to minimize tagging-induced stress. In total, 72 adult winter flounder (27-48 cm) were fitted externally with acoustic transmitters (model V92L, 69 kHz, 9 mmx21 mm, Vemco Ltd., Halifax, NS, Canada) between November 2007 and May 2009. 2 Iwanowicz, H. R., R. D. Anderson, and B. A. Ketschke. 1974. A study of the marine resources of Plymouth, Kingston and Duxbury Bay. Monograph Series Number 17, 37 p. Divi- sion of Marine Fisheries, Boston, MA. Each transmitter had an average pulse rate of 80 sec- onds (range = 40-120 seconds) and an expected battery life of 384 days. A novel external attachment method was developed for this study. Acoustic transmitters were secured into a harness of 9/16 inch (14.3 mm) soft latex tubing by using two-part epoxy (Fig. 2). Two nickel tagging pins were passed upwards from the blind side of the flounder through the dorsal musculature. The harness was then secured to the nickel pins on the eyed-side of the fish with plastic earring backings. The tagging procedure took an average of two to three minutes per fish. A laboratory holding study indicated 100% transmitter retention and 100% survival over a seven-month period. A separate holding study conducted at the University of New Hampshire in 2009 indicated that the tag attachment method did not interfere with the swimming or spawning behavior of tagged fish (E. Fairchild, personal commun.3). In November 2007, 24 prespawning winter flounder were tagged in Plymouth Bay, approximately 5 km from the mouth of the Plymouth Estuary (Fig. 1). Three of these winter flounder were observed to be gravid females at the time of release. Repeated efforts were made to capture spawning flounder within the estuary during March, April, and early May of 2008, but we were not able to capture adult fish until late May 2008. Twenty three adult winter flounder were tagged within the Plymouth Estuary between 24 May and 16 June 2008 (Fig. 1). In the second year of the study, an addi- tional 25 winter flounder were tagged in Plymouth Bay between December 2008 and May 2009 (Fig. 1). Eight winter flounder tagged on 8 May 2009 were observed to be gravid females that were in spawning condition. 3 Fairchild, Elizabeth. 2009. Department of Biological Sci- ences, Univ. New Hampshire, Durham, NH 03824. DeCelles and Cadrin: Movement patterns of Pseudopleuronectes americanus in the southern Gulf of Maine 411 Winter flounder tagged between November 2007 and June 2008 (n = 47) were tracked within the Plymouth estuary by using a moored array of 15 wireless receiv- ers (model VR2 and VR2W, Vemco Ltd.; Fig. 1). Six of these receivers (I, J, K, M, N, and O; referred to as the “inner gate”) were configured as a parallel curtain that spanned the mouth of the Plymouth estuary, allowing the movements of fish traveling between Plymouth Bay and Plymouth estuary to be recorded precisely. Parallel curtain arrays are advantageous because they allow the direction of movement for each fish to be obtained ( Heu- pel et al., 2006). Range testing was performed at the mouth of the estuary by placing transmitters on the bot- tom in different locations for ten-minute intervals. After range testing, receivers within the parallel curtain were positioned to ensure that overlap existed between the detection radii of adjacent receivers. Tagged winter flounder were monitored within the upper reaches of the Plymouth estuary by using nine receivers positioned in a nonoverlapping grid array. These receivers were placed within the deeper channels of the Plymouth estuary to maximize their detection radii. During the second year of the study, the receiver array was expanded from 15 to 30 wireless receivers (Fig. 1). Ten receivers (P-Y; referred to as the “outer gate”) were deployed across the mouth of Plymouth Bay, from Gurnet Point in Duxbury southward to Rocky Point in Plymouth. This arrangement allowed us to track the movement of tagged winter flounder between Cape Cod Bay and Plymouth Bay. An ad- ditional five receivers (Z-DD) were placed within the upper reaches of the Plymouth estuary. Expanding the receiver array allowed us to test more directly our hy- potheses about the spawning behavior of the 25 winter flounder that were tagged between 10 December 2008 and 8 May 2009. Before their retrieval in 2009, six of the 30 acoustic receivers were lost, likely because of boat traffic, in- teractions with commercial fishing gear, or because of winter storms. Receivers P and Y were lost from the outer gate between Cape Cod Bay and Plymouth Bay. Receivers I, M, and O were lost from the parallel cur- tain between Plymouth Bay and the Plymouth estuary. Subsequently, receiver G was moved from the “Cowy- ard” and placed adjacent to the position of receiver M to improve the coverage in this area. Receiver H in Plymouth Harbor was also lost, likely because a high volume of boat traffic in the area. Data analysis The movement track of each tagged fish was visualized by using GIS software (ArcMap vers. 9.3, ESRI, Red- lands, CA). Estuarine residence time was calculated as the time elapsed between the first and last detection for each individual within the Plymouth estuary. If a tagged winter flounder was recorded within the estuary on more than one occasion, the total days spent within the estuary during each visit were combined to calculate residence time for that individual. Winter flounder were classified as either estuarine or coastal spawners on the basis of their observed lo- cations during the peak spawning months of March through May. Fish that were detected within the Plym- outh estuary at any time over this three month span were classified as estuarine spawners, whereas winter flounder that remained in coastal waters were classified as coastal spawners. A G-test for independence (Sokal and Rohlf, 2001) was performed to examine whether significant interannual variability was present in the proportion of tagged winter flounder that were classi- fied as estuarine spawners in 2008 and 2009. The ratio (with 95% confidence intervals) of estuarine to coastal spawners observed over the two-year period was calcu- lated (Sokal and Rohlf, 2001). Abiotic monitoring During the study, bottom water temperature was moni- tored at four locations within the study site. Three temperature loggers (Vemco Ltd.; 8-bit minilog TR) were attached to the moorings of receivers A, E, and K within the Plymouth Estuary. These temperature loggers were placed 0.5 meters above the bottom and were programmed to record bottom water temperature every 30 minutes. A temperature logger placed in the southeastern portion of Plymouth Bay recorded bottom water temperature once every two hours in 2007 and 2008. In September 2008, another temperature logger was placed on the mooring of receiver T to record bottom temperature in Plymouth Bay. In 2009 some of the tem- perature loggers failed. When available, bottom water temperature data were averaged weekly at each site. Results With passive acoustic telemetry, we successfully gath- ered high-resolution data on the migration, spawn- ing behavior, and seasonal distribution of adult winter flounder in the southern Gulf of Maine. Twenty eight of the 47 winter flounder tagged during the first year of the study were later detected within the Plymouth estu- ary, yielding a total of 16,956 detections in our array of acoustic receivers. Twenty two of the 25 fish tagged during the second year of the study were later detected in the receiver array, resulting in 97,394 unique detec- tions. A summary of data for each tagged winter flounder that was later detected is given in Table 1. Spawning behavior Based on the observed movements of tagged winter flounder during the spawning season, the presence of two contingent spawning groups of flounder was evident in the region: coastal spawners and estuarine spawners. In both years of the study, the majority of tagged winter flounder exhibited coastal spawning behavior. Only five of the 24 winter flounder (21%) tagged in Plymouth Bay in 2007 were classified as estuarine spawners. Seven of 412 Fishery Bulletin 108(4) Table 1 Summary of information for tagged winter flounder ( Pseudopleuronectes americanus ) that was detected during the study. For each tagged winter flounder, the date of release and date of last location are given. The total number of detections, and receivers at which that winter flounder was detected, is provided. For the release locations, PB=Plymouth Bay and PE=Plymouth Estuary. Tag ID Length (cm) Release date Release location Last detection No. of detections Receiver stations 5837 34 11/30/2007 PB 8/26/2008 27 Outer Cape 5839 28 11/30/2007 PB 10/30/2008 63 C,F,G,N 5850 30 11/30/2007 PB 11/19/2008 2030 C,J,K,M,N,0 5852 32 11/30/2007 PB 6/1/2008 735 J,K,N,0 5862 29 11/30/2007 PB 6/12/2008 100 C,D,K,M,0 5867 36 11/30/2007 PB 5/28/2008 1 K 5868 34 11/30/2007 PB 8/25/2008 1458 J,K,M,N,0 5877 32 11/30/2007 PB 6/22/2008 799 E,F,G,I,M,0 5832 29 5/24/2008 PE 6/17/2008 44 B,C,J 5833 27 5/24/2008 PE 9/30/2008 3227 B,C,J,K,N 5840 33 5/24/2008 PE 6/10/2008 1681 C,J 5844 27 5/24/2008 PE 6/12/2008 957 C,J,K,M,0 5845 35 5/24/2008 PE 6/13/2008 338 C,J,N 5847 42 5/24/2008 PE 7/30/2008 3198 C,J,K,M,N,0 5855 36 5/24/2008 PE 5/26/2008 90 D 5858 29 5/24/2008 PE 6/15/2008 30 K,M 5859 35 5/24/2008 PE 7/7/2008 18 C,J,N 5866 42 5/24/2008 PE 6/15/2008 119 A,C,J 5873 27 5/24/2008 PE 6/13/2008 267 B,K,M 5878 35 5/24/2008 PE 6/13/2008 217 B,C,J 5835 39 6/6/2008 PE 6/10/2008 55 J,K,M,0 5846 30 6/6/2008 PE 6/21/2008 18 C,J 5856 33 6/6/2008 PE 8/18/2008 211 C,J 5864 27 6/6/2008 PE 6/23/2008 27 B,I,M 5874 38 6/6/2008 PE 6/15/2008 636 B,C,J,N 5871 27 6/9/2008 PE 6/14/2008 9 M,0 5876 36 6/9/2008 PE 9/14/2008 546 C,J,K,N,0 5843 28 6/16/2008 PE 10/6/2008 18 C,K 5872 27 6/16/2008 PE 7/24/2008 37 C,N 53876 33 12/10/2008 PB 5/2/2009 10346 G,K,N,X 53877 31 12/10/2008 PB 8/26/2009 6750 G , Q , R ,T,U,V, W, X , MA Bay 53879 36 12/10/2008 PB 7/25/2009 19394 V,W,X 53884 36 12/10/2008 PB 5/13/2009 1484 Q,R,S,T,U,V,W,X 53886 31 12/10/2008 PB 7/8/2009 14089 G,K,N,S,T,U 53887 30 12/10/2008 PB 7/20/2009 5061 Q,R,S,T,U,V,W,X 53888 31 12/10/2008 PB 6/14/2009 23842 G,J,K,N,R,S,T,U 53890 29 12/10/2008 PB 6/27/2009 57 X 53891 32 12/10/2008 PB 5/22/2009 443 G,K,N,R,S,T,U.V 53894 41 12/10/2008 PB 7/28/2009 3611 C,E,Q,R,S,T,U,V,W,X 53896 31 12/10/2008 PB 6/26/2009 361 X 53892 37 3/10/2009 PB 6/18/2009 1215 R,U,V,W,X 53893 34 3/10/2009 PB 7/8/2009 288 X 53874 33 4/9/2009 PB 4/11/2009 1061 R,S,T,U 53883 35 4/9/2009 PB 5/4/2009 3018 Q,R,S 53885 31 4/9/2009 PB 7/14/2009 2074 G,K,J,N,Q 53873 39 5/8/2009 PB 5/24/2009 620 Q,U,S,T,U,V,W,X 53875 37 5/8/2009 PB 5/12/2009 177 Q,R,S,T,U,V,W 53878 36 5/8/2009 PB 6/12/2009 2230 G,K,Q,R,S,T,U.X 53881 36 5/8/2009 PB 7/25/2009 479 Q,R,S,T,U,V,W,X 53895 36 5/8/2009 PB 6/25/2009 701 B,C,G,J,N,W,X 53897 48 5/8/2009 PB 5/10/2009 93 V,W,X DeCelles and Cadrin: Movement patterns of Pseudopleuronectes americanus in the southern Gulf of Maine 413 T3 O A J L i.i Ial .iiii l f />VV /// ^vv Total number of detections per day C?5 o?5 0*0 rtO o'O nV> AO 00 nO nO" nO nO 00~ nO nO rO ^ ■/ / /- / /■ /• ^ ,# /• & Number of fish detected per day B 1. 1 » _u Al . lift Kt Jm 1 1 ii i ii 8 u A cS? ^ ^ ^ F 4? 4? # jr 4P 4? ^ n# 4$ 44 4P1 4P3 D dp -A3 .rf? AV .Ay o5> 'r'b ■*> $> r5t> rSb rSb kN 3> J 4 4? 4? 12°C. In 2008, 18 of the 23 (78%) winter flounder tagged within the estuary migrated to coastal waters over a three week period from 30 May to 19 June (Fig. 6). This emigration coincided with a sharp increase in water temperature throughout the estuary. As of 19 June 2008, water temperatures exceeded 15°C in all but the deepest portion of the estuary. During July and August of 2008 when temperatures reached their seasonal maxima, no winter flounder were detect- ed in the upper reaches of the estuary, and six winter flounder were detected at the mouth of the estuary, where temperatures were the coolest. In July 2009, only one winter flounder was detected in the estuary, for a period of one day. Our observations on the autumn distribution of win- ter flounder in the Plymouth estuary were limited to 2008. Results indicated that five tagged individuals re- turned to the estuary as water temperatures decreased below their seasonal maxima. For tagged winter flounder that were detected in the estuary, residence times within the estuary ranged from 2 to 91 days (mean=29 days). The estuarine residence times of fish tagged in Plymouth Bay (mean =40 days) were significantly greater than the residence times for winter flounder tagged in the estuary (mean=22 days) (single-factor ANOVA; P=0.025). No re- lationship was found between fish size and estuarine residence time (P=0.88). For winter flounder tagged within the Plymouth estuary in 2008, the observed residence time is almost certainly an underestimate because it is unknown how long these fish were present in the estuary before being tagged. Detections of tagged winter floun- der were not uniformly distributed throughout the study site. In the first year of the study, the inner gate of acoustic receivers typically detected more flounder and had more detections than receivers located within the estu- ary (Fig. 7, A and B). Tagged winter flounder were detected throughout the estuary, and seven of the nine receiv- ers inside the estuary detected at least one tagged flounder. In the second year of the study, tagged winter flounder were most commonly detected in the coastal waters of Plymouth Bay. The outer gate of receivers in Plymouth Bay typically detected the greatest number of winter flounder and had the larg- est number of detections (Fig. 7, C and D). Two tagged winter flounder (Tags 53876 and 53886) remained at the inner gate for long periods of time within the detection radii of receivers G and K, which accounted for the large number of detections by these receiv- ers (Fig. 7D). In 2009, tagged winter flounder were detected at only three of the 12 receivers inside the estuary, and the low number of detections (<1000 ) at these three receivers indicates that residence within the estuary was brief. The bimodal distribution of receiver detections (e.g., Fig. 70 supports the inference of divergent habitat use, spawning behaviors, and contingent structure. Two tagged winter flounder were detected at receivers used by other Total number of detections per day Vi6 X? X? X? aC? X? aC? ^ x? ^ x? x? ^ T> T>1' " vv -->■ Number of fish detected per day ■o nj 2 ^ XS° a<5° VT aO3 nX5° XT’ rS>J .XT aO Percentage of receivers with detections, per day Figure 4 Temporal patterns of winter flounder ( Pseudopleuronectes americanus ) detections obtained by using passive acoustic telemetry in the Plym- outh Estuary and Plymouth Bay, MA, between December 2008 and July 2009. (A) The total number of detections each day. (B) The number of tagged winter flounder that were detected each day. (C) The percentage of acoustic receivers that detected a tagged winter flounder during each day of the study. DeCelles and Cadrin: Movement patterns of Pseudopleuronectes americanus in the southern Gulf of Maine 415 O J5P /■ ^ ^ xT Figure 5 Weekly averaged bottom water temperatures observed between November 2007 and July 2009. Water temperature was recorded every 30 minutes at four locations during the study: Plymouth Bay; the mouth of the Plymouth Estuary; Kingston Bay (western portion of Plymouth Estuary); and Duxbury Bay (northern portion of Plymouth Estuary). O 2008 Figure 6 The relationship between water temperature and the emigration of tagged winter flounder (Pseudopleuronectes americanus ) from the Plymouth Estu- ary in 2008. Winter flounder emigrated rapidly from the Plymouth Estuary as water temperatures increased from 10° to 15°C at the beginning of the summer. researchers (W. Hoffman, personal commun.4), documenting more sub- stantial movements beyond the study area. One tagged winter floun- der (tag 5837) moved from the Gulf of Maine stock area to the SNE-MA stock area and was detected east of Cape Cod on 26 August 2008 (Table 1). This individual was tagged in Plymouth Bay on 30 November 2007 and was never detected in our re- ceiver array. Tag 53877 was detected throughout our array in 2009 and was later detected in Massachusetts Bay on 26 August 2009 (Table 1). Receiver efficiency Despite the loss of some receivers, it appears that the array was efficient for the detection of the migrations of winter flounder. Twenty eight winter flounder released during the first year of the study were later detected in the estuary. Each fish was detected in the estuary an aver- age of 416 times (range = l-3227 detections) at an average rate of 1.9 detections/day at liberty. Seventeen of the 24 winter flounder released in Plymouth Bay were never detected inside the estuary. These 24 winter flounder were released outside the receiver array that was used in the first year (Fig. 1); therefore, if these fish remained in coastal waters they would not have been detected by our array. Twenty two of the 25 winter flounder released in the second year of the study were later detected in our array. On average, each tagged winter flounder was detected 2861 times (range 57-23,842 detections) at an average rate of 14.1 detec- tions/day at liberty. Winter flounder tagged on 8 May 2009, were released slightly eastward of the receiver array (Fig. 1), and this location may explain why two of these tagged fish were never detected again. In the fall of 2008, receivers I and M were lost from the inner gate of receivers that spanned the mouth of the estuary. After these receivers were lost, two tagged winter flounder (tags 5833 and 5873) were detected in- side the estuary at the end of September and October, respectively, without being detected at the inner gate of receivers. One flounder (tag 5855) was released inside 4 Hoffman, William. 2009. Massachusetts Division of Marine Fisheries, Gloucester, MA 01930. the estuary on 24 May 2008, and was detected two days later at receiver D. However, this individual was never detected at another receiver during the study. The fate of this fish is unknown. It may have been captured in the recreational fishery, eaten by another animal, or died after having been tagged. This was the only win- ter flounder that was released and detected inside the estuary (tag 5855) that was never detected at the inner gate. In the second year of the study, only one tagged winter flounder (tag 53894) that was detected inside 416 Fishery Bulletin 108(4) a3 Aouenbejj m CJ •2 ^ oi ^ 0) o o ° > O GO o o 53 cm S (D * a T3 (D SPQ C £ 03 3 03 » oo o o CM ^ -Q ^ S -2 ° ^ > a o C/3 8 yr; Fig. 3). The maximum age in the catch also 424 Fishery Bulletin 108(4) differed among species (Fig. 3) and the oldest fish was a 36-year-old S. nematophorus. Regional and sex-specific differences in length- and age-frequency distributions could only be analyzed for L. carponotatus because of low sample sizes for all other species. There was a significant effect of sex on lengths (Kolmogorov-Smirnov test: n1=1279, n2= 399, dmax=0-336, D = 0.078, P<0.05) and ages (n1 = 1279, n2- 399, c?max=0.281, .0=0.078, P<0.05) for the frequency distributions and a greater proportion of males than fe- males in the larger size classes and younger age classes (Fig. 4). Differences in growth by sex were significantly different (^2 = 68.34, PcO.OOOl), with males reaching a significantly greater maximum size than females (female: L^- 278, ,K=0.43, t0=-2.55; male: 0^=293, K= 0.55, f0=-1.43). The unconstrained fits of the VBGF differed substan- tially among species with L. carponotatus, A. virescens, and L. fulviflamma reaching a maximum size relatively early in life (Fig. 5). Symphorus nematophorus reached a maximum size somewhat later in life, and the growth curve for L. gibbus was not asymptotic. The lack of juvenile fish collected for all species, but particularly for L. carponotatus, L. fulviflamma, and L. gibbus, re- sulted in relatively flat fits of the unconstrained VBGF. The nonasymptotic growth pattern for L. gibbus indi- cates that under-sampling of larger individuals may also have occurred. VBGF parameter estimates for all species from the unconstrained fit are likely to be biased and should be interpreted with some caution. Constraining the VBGF by setting t0=0 produced similar estimates of K and Lx to those for the unconstrained estimates for L. carponotatus and L. ful- viflamma. Despite limited sam- pling of the youngest individu- als, this similarity indicates that a sufficiently wide range of age classes were sampled in these populations to produce bi- ologically reasonable estimates of growth without the need to constrain tQ. Constraining t0 re- sulted in faster initial growth estimates (greater estimate of K) and smaller asymptotic length (LJ than the uncon- strained fits for S. nematopho- rus, A. virescens, and L. gibbus. The constrained fit produced a more asymptotic growth curve for L. gibbus, indicating that constraining t0 may have pro- vided a more biologically realis- tic estimate of length-at-age for this species. Both constrained and uncon- strained growth curves differed across species. These species showed distinct differences in the rate of growth at young ages and the age at which they reached average maxi- mum length: smaller species grew fast in the first couple of years and reached asymptotic length early, whereas larger species grew slightly slower in the first few years to reach an asymptotic FL later. The lack of small-size individuals resulted S. nematophorus o= 168 i£U ri n~HT-n ,[_ixi □ , n n n ■(NICMCMCNCNIfOfOfOCO ill L gibbus a=166 10 L fulviflamma o=55 IQ Age Figure 3 Age-frequency distributions for five lutjanid species on the Great Barrier Reef between 1995 and 2005 (note difference in y axes). Heupel et al. : Demographic characteristics of exploited tropical lut|anids 425 in a lack of definition of early growth in the un- constrained curves when compared to constrained estimates. Both constrained and unconstrained growth curves for L. carponotatus and L. fulvi- flamma were relatively “square-shaped” (Choat and Robertson, 2002), revealing little growth over most of their life spans. Several species had K values >0.30/yr, indicating rapid growth to maximum size at a young age. Comparison of growth rate among 5-12 year olds revealed significant differences among species (F4 2503=1543.8, PcO.OOOl). Post hoc tests indicated that each species had a unique growth rate. Exami- nation of age at 50% Lx indicated some differences between constrained and unconstrained fits (Table 1), but in all cases, 50% of asymptotic length was reached at relatively young ages compared with estimated longevity, and only L. gibbus took longer than 10% of expected longevity to reach 50% of asymptotic size (Table 1). Lutjanus carponotatus was the only species with a sample size sufficient for regional analysis of growth data. Likelihood ratio tests indicated that patterns of growth differed significantly between Lizard Island, Mackay, and Storm Cay regions (^2 = 68.34, P<0.001) but there were similar Lx values among regions, and quite variable K val- ues (Table 1). The most notable difference was the greater (and lower K) in the Storm Cay region (Table 1). Total mortality (Z) estimates calculated from catch curves varied significantly among the five species examined (F441=5.61, P=0.001) (Fig. 6). Z was highest for L. gibbus and A. virescens and lowest for L. fulviflamma (Table 2). Mortality rates calculated by the Hoenig method resulted in more homogeneous estimates, of which highest rates were for L. gibbus and lowest for S. nematophorus. Mor- tality estimates calculated from catch curves were higher than those estimated by Hoenig’s method for all species except L. fulviflamma (Table 2). Z for L. carponotatus did not vary significantly among the three regions (P2 29 = 2.24, P=0.15) and therefore a single catch curve was fitted to data pooled across regions. All except one of the individuals sampled (282-mm L. gibbus ) were sexually mature, making analysis of size and age at maturity impossible. Adult sex ratios in a few of the sampled populations differed from a 1:1 sex ratio (Table 3), despite the expected gonochoristic nature of these species. Sex ratios of L. carponotatus, L. gibbus, and L. vitta were biased significantly toward males. Discussion Many coral reef fish species have overlapping spatial distributions, are found in similar habitats, and grow to roughly similar sizes. It might be expected that superfi- cially similar species from the same family would have similar demographic characteristics and life history strategies. The results of this case study of lutjanid species, however, indicate that it is impossible to make generalizations about the life histories of members of a family, even where they grow to similar sizes and coexist in the same habitats. Even this small subset of seven lutjanid species had quite different growth, mortality, and longevity characteristics. Our results support pre- vious findings for individual species but also provide evidence of the considerable differences among species in this family and highlight the need for careful, and potentially species-specific, approaches to managing the harvest of lutjanid assemblages. Demographic characteristics of the lutjanid species described here must be considered in the context of the limitations of size selectivity of the fishing gear used to sample populations. No individuals below 200 mm FL were collected for any species, meaning the youngest age classes were not fully selected, resulting 426 Fishery Bulletin 108(4) in an absence of immature individuals for all, but one species. However, some individuals of most species were sampled in their first or second year indicating that these species mature at a very early age. The fact that full recruitment to the gear occurred at age 2 for S. nematophorus and at age 3 for A. virescens, the larg- est species sampled, illustrated the rapid early growth characteristic of many lutjanid species (Newman et al., 1996, 2000a; Kritzer, 2004; Grandcourt et al., 2006). It is important to note that the same sampling gear was used in this project as that most commonly used by commercial and recreational fisheries on the GBR and therefore the biological characteristics described here should reflect those characteristics of individuals of these species harvested by the fishery. Importantly, the sampled age distributions revealed the age range of these populations susceptible to fishing gear. It is also likely that many fish were not sampled because sampling was constrained to depths shallower than 30 m and to daylight hours, which are the most common Heupel et al.: Demographic characteristics of exploited tropical lutjanids 427 Table 1 Parameter estimates for the von Bertalanffy growth function parameters age at 50% (mean asymptotic fork length) and per- cent longevity at 50% Lx for five lutjanid species from the Great Barrier Reef. K is the von Bertalanffy growth coefficient, and t0 is the theoretical age at length zero. “U” indicates an unconstrained estimate, “C” indicates an estimate constrained by t0= 0. Region-specific growth curves for L. carponotatus were not constrained. Species L ^ (mm) K (/yr) t0{ yr) Age (yr) at 50% % longevity at 50% S. nematophorus — U 820 0.12 -2.83 3 8.33 S. nematophorus — C 732 0.26 0 3 8.33 A. virescens — U 683 0.35 -1.53 0.5 3.13 A. virescens — C 623 0.85 0 1 6.25 L. gibbus — U 544 0.06 -9.48 2 16.67 L. gibbus — C 352 0.51 0 3 25.0 L. fulviflamma — U 265 0.61 1.66 1 5.88 L. fulviflamma — C 267 0.41 0 1 5.88 L. carponotatus — U 295 0.37 -2.58 0 0 L. carponotatus — C 291 0.66 0 1 4.35 L. carponotatus — Lizard Is 293 0.67 -0.17 — — L. carponotatus — Mackay 292 0.47 -1.68 — — L. carponotatus — Storm Cay 302 0.16 -9.35 - — — Estimates of mortality for five lutjanid species using catch used in mortality estimation are indicated. Table 2 curve and Hoenig (1983) estimators. Maximum age and age range Species Maximum age (yr) Age range (yr) Catch curve Z (/yr) [SE] Hoenig Z (/yr) S. nematophorus 36 3-17 0.20 [0.05] 0.11 A. virescens 16 2-9 0.56 [0.07] 0.26 L. gibbus 12 8-12 0.63 [0.08] 0.35 L. carponotatus 23 7-23 0.30 [0.02] 0.18 L. fulviflamma 17 10-17 0.14 [0.06] 0.25 depth and time fished by commercial and recreational fishing crews. A small number of fishing crews fish dur- ing the night in deeper water (> 50 m) where a greater proportion of larger lutjanids are caught (Marriott, personal commun.). Limited sample sizes prevented the statistical analy- ses of sex- and region-specific variation in life history characteristics except for L. carponotatus. As with pre- vious findings, L. carponotatus displayed sex-specific differences in these characteristics, with males reach- ing larger maximum sizes than females. Reasons for sex-specific size and age distributions may be due to a wide array of factors, including regional conditions and physiological costs of sperm and egg production. Kritzer (2004) also found male L. carponotatus to be larger than females and suggested that this is an Indo-Pacific trait within the lutjanids because it does not always occur in other geographic areas. Such sex-specific differ- ences in lutjanid populations also may vary by region, although the magnitude of any sex-dependent variation within a geographic area remains poorly understood. Hence, our parameter estimates for the remaining spe- cies from samples pooled over regions of the GBR may not be as informative as we would like, but they nev- ertheless provide a good starting point for comparison and applicability to other regions where data remain scarce or absent. Age and growth Five of the seven species examined grew to similar sizes, but three of these demonstrated different longevities and the two largest species showed a twofold range in longevity for roughly similar maximum sizes, again illustrating the lack of a relationship between length and age across tropical lutjanids. High maximum ages 428 Fishery Bulletin 108(4) Table 3 Number and size of male and female lutjanid individuals sampled from the Great Barrier Reef between 1995 and 20o5 for re- productive analysis. Size range is included in parentheses; values in bold indicate significant difference from an even sex ratio. Species Female Male Sex ratio F/M n Mean size (mm FL) n Mean size (mm FL) L. fulviflamma 28 263 16 250 1.75 (209-315) (219-285) S. nematophorus 32 533 41 485 0.78 (340-772) (320-789) A. virescens 50 511 32 509 1.56 (332-810) (334-778) L. adetii 18 284 22 290 0.82 (254-374) (243-341) L. gibbus 8 277 48 330 0.17 (227-356) (262-418) L. vitta 7 266 18 270 0.39 (204-307) (226-310) L. cai'ponotatus 520 274 1722 292 0.30 (203-355) (205-405) of 36 and 23 years were evident in S. nematophorus (the largest species) and L. carponotatus (the second smallest species), respectively, despite their different maximum sizes. Lutjanus gibbus, A. virescens, and L. fulviflamma had shorter life spans and maximum ages of 12, 16, and 17 years, respectively. Lutjanus carponotatus have previ- ously been reported to have maximum ages of 18 and 20 years (Newman et al., 2000a; Kritzer, 2002, 2004) which are consistent with the 23 year maximum obtained here. Shorter longevities recorded for L. gibbus, A. virescens, and L. fulviflamma align with those from other lutja- nid species such as L. vitta (12 years, Newman et al., 2000a), L. guttatus (rose snapper, 11 years, Amezcua et al., 2006), and L. fulviflamma (14 years, Grandcourt et al., 2006). In contrast, L. adetii, a small species simi- lar in size and appearance to L. vitta, has a reported maximum age of 24 years (Newman et al., 1996), further demonstrating the variability in life histories of similar size lutjanid species. This and most previous studies reveal that these species have rapid initial growth despite differences in size and longevity; however, they reach asymptotic sizes at relatively young ages (Newman et al., 1996, 2000a; Kritzer, 2002, 2004; Grandcourt et al., 2006). Newman et al. (1996) suggested that size could not be used to infer age for L. adetii beyond five years of age. This principle is probably applicable for many lutjanid species and is supported by the growth patterns for four of the five species for which growth was analyzed, the exception being L. gibbus. All five species reached 50% of before reaching 4 years of age. Caution should be used when generalizing, however, because red bass (L. bohar) and goldband snapper ( Pristipomoides multidens) have been shown to be slow growing, long lived, and late maturing, showing a classic K-selected life history (Newman and Dunk, 2003; Marriott et al., 2007). That result contrasts with the fast growth observed in most other lutjanid species so far examined. Lutjanus gibbus was the only species that did not reach an asymptote in the unconstrained growth curve. This may have been the result of the gear excluding smaller and younger individuals, a proposition support- ed by the fact that the constrained VBGF curves did produce an asymptote and the unconstrained functions produced an unrealistic t0 value of -9.48. Observations that larger individuals were collected from deeper water indicate that the largest size and oldest age classes of L. gibbus and S. nematophorus may have been under- sampled. Alternatively, these species may indeed have a long growth period and not stop growing throughout life, as does L. bohar (Marriott et al., 2007). The larg- est reported L. gibbus, however, was 50 cm total length (TL) (Carpenter and Niem, 2001) — only 5 cm larger than the largest individual sampled in this study (418 mm FL, =500 mm TL). Lutjanus gibbus was also the species with the highest mortality estimates from both catch curve and Hoenig’s estimators, indicating that it had the highest turnover rates. Further sampling will be required to gain a better understanding of the age and growth characteristics of this species and resolve whether the differences in characteristics between L. gibbus and the other species examined were attribut- able to different life history strategies or to the effects of species-specific sampling biases. All five species for which age was estimated reached 50% of the estimated L „ at young ages in relation to their maximum age. The percentage of longevity at which 50% was reached was similar for most spe- Heupel et al.: Demographic characteristics of exploited tropical lutjanids 429 4 5 3 • • • S. nematophorus 4 \ * A. virescens 2 1 3 • \ X 2 • • • • X • • i • \ • X* 0 Xv 2 4 6 8 10 12 14 16 18 20 0 2 4 6 8 10 12 8 6 In frequency hO •&. O) \ 5 L. carponotatus • 4 * . • 2 • • 1 • L. gibbus • X. 0 o 8 10 12 14 16 18 20 22 24 6 8 10 12 3 L. fulviflamma 2 x. • 1 • • Xs^ • 0 • • 5 8 10 12 14 16 18 20 Age (yr) Figure 6 Catch curves for five lutjanid species from the Great Barrier Reef between 1995 and 2005. The slopes of the regressions provided estimates of the rate of total mortality ( Z ) for each species. cies, and L. gibbus was the only species taking greater than 10% of its longevity to attain 50% Lm. Growth of L. carponotatus varied significantly among regions spanning 7° latitude, growing more quickly to smaller asymptotic lengths in the more equatorial regions than in the south. Although statistically significant, the dif- ference in growth may have little biological signifi- cance with varying only about 3% among regions. Regional variation in growth for the other species re- mains unknown because of a lack of data but results for L. carponotatus reveal potential environmental influ- ences on observed patterns for these other species. Mortality Estimates of mortality were likely biased because of selectivity of the sampling gear. For example, the posi- tively skewed age distributions for S. nematophorus and A. virescetrs indicate that older fish were less avail- able to the sampling gear and were likely to be under- 430 Fishery Bulletin 108(4) represented in the sample, resulting in overestimates of Z from catch curves. Estimates of Z derived from Hoenig’s equation may be more appropriate for these species. Similarly, older individual L. gibbus and L. fulviflamma may not have been collected by the gear and if so, under-sampling may have resulted in overes- timates of Z from the catch curves and Hoenig’s method. Notwithstanding these potential biases, comparison of mortality rates revealed that A. virescens had a similar mortality rate to that of L. gibbus, indicating that A. virescens also has a relatively high population turnover rate. Aprion virescens grows quickly to a large size but may not live as long as other lutjanid species, despite attaining relatively large sizes. Symphorus nematophorus, for example, attained the largest size and had the second lowest mortality rate of the species examined. Lutjanus fulviflamma mortality estimates differed most between estimation methods, with catch curve estimates resulting in lowest mortality rates for this species (0.14/yr), but the Hoenig estimate revealed higher mortality rates (0.25/yr), similar to those for A. virescens and L. gibbus. The Hoenig estimate was also similar to rates previously reported for L. fulvi- flamma (0.29/yr, Grandcourt et ah, 2006), indicating some bias in the estimate from catch curves. Like L. gibbus, L. carponotatus under six years of age could not be used in catch curve estimates of mortality, but the 6-23 year age classes yielded mortality estimates for L. carponotatus of 0.30/yr (catch curve) and 0.18/yr (Hoenig), similar to estimates calculated previously for this species (0.20/yr, Newman et al., 2000a; 0.26-0.29/ yr, Kritzer, 2004). The range of mortality estimates for the species examined in this analysis agree well with those for other lutjanids (Newman et ah, 1996, 2000a; Amezcua et al., 2006) and are another indication of the variability in life history strategies within the family. Maturity and sex ratio All except one individual sampled across all species were sexually mature, supporting the conclusion that these lutjanids reach sexual maturity early in life. Biased sex ratios were observed for all of the species sampled although the apparent biases were statistically significant for only three species. Two species (L. fulvi- flamma, A. virescens ) showed large but nonsignificant female-biased sex ratios, whereas all others showed a male-biased ratio, and three (L. carponotatus, L. vitta, L. gibbus) were not significantly different from 1:1. Lutjanids are gonochoristic species and therefore it may be expected that adult sex ratios would be close to 1:1 in local populations, although at least three other studies have revealed biased sex ratios. Kritzer (2004) found that L. carponotatus had a female-biased sex ratio, whereas Newman et al. (2000a) reported a strongly male-biased sex ratio for the same species in similar locations to those that we examined. Studies of L. fulviflamma from different locations showed widely variable sex ratios. Kaunda-Arara and Ntiba (1997) reported a male-biased sex ratio in Kenya, and Grand- court et al. (2006) reported a female-biased sex ratio in the southern Arabian Gulf. It is difficult, therefore, to establish generic patterns of sex ratios across, or even within, lutjanid species given the contradictory patterns in the literature and among the species sampled here. It is possible that sex-ratio bias is a result of a differential survival of males and females or sex-specific patterns in distribution that would result in males and females having different probabilities of capture in the sampling strategies used in various studies, including the present one. Kritzer (2004) was one of the few to have examined mortality by sex, but no difference in mortality by sex was found. This single result may indicate that differ- ential spatial distributions may be a more likely cause of sex-ratio biases in samples of lutjanid populations. It has been suggested that sex ratio may be more even during spawning events (Kritzer, 2004), but no data are available to test this hypothesis. Implications for management of lutjanid populations An improved understanding of the demographic param- eters of gonochoristic species is crucial to furthering research on the effects of fishing on their populations. Results here clearly demonstrate that all or even super- ficially similar (e.g., in size) subsets of lutjanid species should not be treated in the same manner in these types of analyses and that vulnerability is likely to be variable within the family. Careful consideration of the inherent life history variability of these species is required in the development of theories and generalizations about this family and others. This case study of seven lutjanid species clearly in- dicates that data from one species cannot be applied to another in determining appropriate management measures for these populations. Some species appeared to be more susceptible than others to overexploitation by fisheries because of their longer life spans and lower rates of mortality. Symphoi'us nematophorus is a good example of a species with a long life span, large size, and low mortality. The combination of these factors could make this species a desirable fisheries target (owing to its larger size than other lutjanids), but also one of the most vulnerable because of its life history characteristics. Symphorus nematophorus and L. gibbus are currently no-take species within the GBR and have historically been avoided because of a potential risk of ciguatera (a form of food poisoning from eating large reef fish), and therefore these species are largely pro- tected from harvest within this region, although this may not be the case in other parts of their range. In comparison, the longevity of L. carponotatus and L. adetii in relation to other lutjanid species may make them more vulnerable to overfishing than sibling spe- cies (Newman et al., 1996). Despite their small size and fast growth, these populations may be more vulnerable than similar species such as L. vitta which have shorter longevities and hence potentially faster population turn- over rates (Newman et al., 2000a). Notably, most of the species examined here recruited to the fishing gear at Heupel et al.: Demographic characteristics of exploited tropical lutjamds 431 relatively early ages but also apparently reached sexual maturity before recruitment to the fishery. The exis- tence of early maturity in relation to longevity indicates a life history strategy with high reproductive potential in older individuals — perhaps indicative of adaption to low-frequency episodic recruitment successes (Kritzer, 2002, 2004). These characteristics may mean that lu- tjanids are particularly vulnerable to recruitment col- lapse from sustained harvest at relatively low rates of fishing-induced mortality. Current management measures for lutjanid species within the GBR are precautionary to protect the wid- est possible range of lutjanid family members, but this may not be the case in other parts of their range and implications of differences among species should be con- sidered in such situations. The presence of substantial variation in life histories among species would indicate that the data presented here could serve as a bench- mark for these species in other regions, but should not be applied to different species that may not have com- parable demographic parameters despite morphometric similarities to the species examined here. Failure to recognise differences in life histories when implement- ing management strategies across species may result in over-exploitation of some species, under-exploitation of others, or both. Furthermore, we recommend that local species-specific estimates of population parameters are obtained wherever possible because significant regional and local variation in population parameters is becom- ing increasingly apparent for tropical reef fish popula- tions (e.g., Adams et al., 2000; Kritzer, 2002; Williams et al., 2003). Consequently, estimates derived from local populations are likely to significantly improve assess- ments and advice for the management of particular stocks. Acknowledgments Funding for the ELF experiment was provided by the CRC Reef Research Centre, the Fisheries Research and Development Corporation, the Great Barrier Reef Marine Park Authority, Queensland Primary Industries and Fisheries, and James Cook University. The authors would like to thank the many commercial fishermen who participated in the ELF experiment, in particular R. Stewart, M. Petersen, T. Must, K. Holland, and M. Bush. Funding to process and analyze samples from the ELF experiment was provided by the Australian Government’s Marine and Tropical Sciences Research Facility through the Reef and Rainforest Research Centre. This manuscript is a contribution from the ELF project. Literature cited Adams, S. 2003. Morphological ontogeny of the gonad of three plectropomid species through sex differentiation and transition. J. Fish Biol. 63:22-36. Adams S, B. D. Mapstone, G. R. Russ, and C. R. Davies 2000. Geographic variation in the sex ratio, sex spe- cific size, and age structure of Plectropomus leopardus (Serranidae) between reefs open and closed to fishing on the Great Barrier Reef. Can. J. Fish. Aquat. Sci. 57:1448-1458. Amezcua, F., C. Soto-Avila, and Y. Green-Ruiz. 2006. Age, growth, and mortality of the spotted rose snapper Lutjanus guttatus from the southeastern Gulf of California. Fish. Res. 77:293-300. Arreguin-Sanchez, F., and S. Manickchand-Heileman. 1998. The trophic role of lutjanid fish and impacts of their fisheries in two ecosystems in the Gulf of Mexico. J. Fish Biol. 53 (Suppl. A):143-153. Campbell, R. A., B. D. Mapstone, and A. D. M. Smith. 2001. Evaluating large-scale experimental designs for management of coral trout on the Great Barrier Reef. Ecological Applications 11:1763-1777: Appendix A — the REEF simulation model and appendix B — statistical models and analyses. Ecol. Archive A011-021. Cappo, M., R Eden, S. J. Newman, and S. Robertson 2000. A new approach to validation of periodicity and timing of opaque zone formation in the otoliths of eleven species of Lutjanus from the central Great Barrier Reef. Fish. Bull. 98:474-488. Carpenter, K. E., and V. H. Niem. 2001. FAO species identification guide for fishery pur- poses. The living marine resources of the Western Central Pacific, vol. 5, Bony fishes part 3 (Menidae to Pomacentridae), p. 2791-3379. FAO, Rome. Choat, J. H., J. P. Kritzer, and D. R. Robertson. 2009. Ageing in coral reef fishes: Do we need to validate the periodicity of increment formation for every spe- cies of fish for which we collect age-based demographic data? In Tropical fish otoliths: information for assess- ment, management and ecology, reviews: methods and technologies in fish biology and fisheries, vol. 11 (B. S. Green, B. D. Mapstone, G. Carlos, and G. A. Begg, eds.). Springer, Netherlands, doi 10.1007/978-1-4020- 5775-5_2. Choat, J. H., and D. R. Robertson. 2002. Age based studies. In Coral reef fishes (P. F. Sale, ed.), 57-80. Academic Press, San Diego, CA. Davis, T. L. O., and G. J. West. 1992. Growth and mortality of Lutjanus vittus (Quoy and Gaimard) from the north west shelf of Australia. Fish. Bull. 90:395-404. Fowler, A. J. 1995. Annulus formation in coral reef fish — a review. In Recent developments in fish otolith research (D. H. Secor, J. M. Dean and S. E. Campana, eds.), p. 4563. Univ. South Carolina Press, Columbia. SC. Ferreira, B. P. 1995. Reproduction of the common coral trout Plectropo- mus leopardus (Serranidae: Epinephlinae) from the cen- tral and northern Great Barrier Reef, Australia. Bull. Mar. Sci. 56:653-669. Grandcourt, E. M., T. Z. Al Abdessalaam, and F. Francis. 2006. Age, growth, mortality and reproduction of the blackspot snapper, Lutjanus fulviflamma (Forsskal, 1775), in the southern Arabian Gulf. Fish. Res. 78:203-210. Gust, N., J. Choat, and J. Ackerman. 2002. Demographic plasticity in tropical reef fishes. Mar. Biol. 140:1039-1051. 432 Fishery Bulletin 108(4) Hoenig, J. M. 1983. Empirical use of longevity data to estimate mor- tality rates. Fish. Bull. 82:898-902. Jennings, S., J. D. Reynolds, and S. C. Mills. 1998. Life history correlates of responses to fisheries exploitation. Proc. Royal Soc. Lon. Ser. B, Biol Sci 265:333-339. Kamukuru, A. T., T. Hecht, and Y. D. Mgaya. 2005. Effects of exploitation on age, growth and mortality of the blackspot snapper, Lutjanus fulviflamma, at Mafia Island, Tanzania. Fish. Manag. Ecol. 12:45-55. Kaunda-Arara, B., and M. J. Ntiba. 1997. The reproductive biology of Lutjanus fulviflamma (Forsskal. 1775) in Kenyan inshore marine waters. Hy- drobiologia 353:153-160. Kritzer, J. P. 2002. Variation in the population biology of stripey bass Lutjanus carponotatus within and between two island groups on the Great Barrier Reef. Mar. Ecol. Prog. Ser. 243:191-207. 2004. Sex-specific growth and mortality, spawning season, and female maturation of the stripey bass ( Lut- janus carponotatus) on the Great Barrier Reef. Fish. Bull. 102:94-107. Marriott, R. J., B. D. Mapstone, and G. A. Begg. 2007. Age-specific demographic parameters, and their implications for management of the red bass, Lutjanus boliar (Forsskal 1775): a large, long-lived reef fish. Fish. Res. 83:204-215. Meyer, C. G., Y. P. Papastamatiou, and K. N. Holland. 2007. Seasonal, diel, and tidal movements of green jobfish ( Aprion virescens , Lutjanidae) at remote Hawaiian atolls: implications for marine protected area design. Mar. Biol. 151:2133-2143. Molloy, P. P, N. B. Goodwin, I. M. Cote, M. J. G. Gage, and J. D. Reynolds. 2007. Predicting the effects of exploitation on male-first sex-changing fish. Anim. Conserv. 10:30-38. Musick, J. A., M. M. Harbin, S. A. Berkeley, G. J. Burgess, A. M. Eklund, L. Findley, R. G. Gilmore, J. T. Golden, D. S. Ha, G. R. Huntsman, J. C. McGovern, S. J. Parker, S. G. Poss, E. Sala, T. W. Schmidt, G. R. Sedberry, H. Weeks, and S. G. Wright. 2000. Marine, estuarine, and diadromous fish stocks at risk of extinction in North America (exclusive of Pacific salmonids). Fisheries 25:6-30. Newman, S. J., M. Cappo, and D. McB. Williams. 2000a. Age, growth and mortality of the stripey, Lutjanus carponotatus (Richardson) and the brown-stripe snap- per, L. vitta (Quoy and Gaimard) from the central Great Barrier Reef, Australia. Fish. Res. 48:263-275. Newman, S. J., M. Cappo, and D. McB. Williams. 2000b. Age, growth, mortality rates and correspond- ing yield estimates using otoliths of the tropical red snappers, Lutjanus erythropterus, L. malabaricus and L. sebae from the central Great Barrier Reef. Fish. Res. 48:1-14. Newman, S. J., and I. J. Dunk. 2003. Age validation, growth, mortality, and additional population parameters of the goldband snapper {Pristi- pomoides multidens) off the Kimberley coast of North- western Australia. Fish. Bull. 101:116-128. Newman, S. J., D. McB. Williams, and G. R. Russ. 1996. Age validation, growth and mortality rates of the tropical snappers (Pisces: Lutjanidae) Lutjanus adetii (Castelnau, 1873) and L. quinquelineatus (Bloch, 1790) from the central Great Barrier Reef, Australia. Mar. Freshw. Res. 47:575-584. Parent, S., and L. M. Schriml. 1995. A model for the determination of fish species at risk based upon life-history traits and ecological data. Can. J. Fish Aquat. Sci. 51:1768-1781. Pilling, G. M., R. S. Millner, M. W. Easey, C. C. Mees, S. Rathacharen, and R. Azemia. 2000. Validation of annual growth increments in the oto- liths of the lethrinid Lethrinus mahsena and the lujanid Aprion virescens from sites in the tropical Indian Ocean, with notes on the nature of growth increments in Pris- tipomoides filamentosus. Fish. Bull. 98:600-611. Polovina, J. J., and S. Ralston. 1987. Tropical snappers and groupers: biology and fisher- ies management, 659 p. Westview Press, Boulder, CO. Ricker, W. E. 1975. Computation and interpretation of biological sta- tistics of fish populations. Fish Res. Board Can. 191, 382 p. Roff, D. A. 1984. The evolution of life-history parameters in tele- osts. Can. J. Fish Aquat. Sci. 41:989-1000. Sadovy, Y., and D. Y. Shapiro. 1987. Criteria for the diagnosis of hermaphroditism in fishes. Copeia 1987:136-156. Sale, P. F. 1991. The ecology of fishes on coral reefs, 754 p. Aca- demic Press, San Diego, CA. Shimose, T., and K. Tachihara. 2005. Age, growth and maturation of the blackspot snapper Lutjanus fulviflamma around Okinawa Island, Japan. Fish. Sci. 71:48-55. Sokal R. R„ and F. J. Rohlf. 1995. Biometry. The principles and practice of statistics in biological research, 3rd ed., 893 p. W. H. Freeman and Co., New York. Williams A. J., C. R. Davies, B. D. Mapstone, and G. R. Russ. 2003. Scales of spatial variation in demography of a large coral reef fish: An exception to the typical model? Fish. Bull. 101:673-683 Wilson, C. A., and D. L. Nieland. 2001. Age and growth of red snapper, Lutjanus campechanus, from the northern Gulf of Mexico off Louisiana. Fish. Bull. 99:653-664. Zar, J. H. 1999. Biostatistical analysis, 4th ed., 929 p. Prentice- Hall Inc., Englewood Cliffs, NJ. 433 Abstract — We evaluated the conser- vation benefits of the use of circle hooks compared with standard J hooks in the recreational fishery for Atlantic istiophorid billfishes, noting hooking location and the presence of trauma (bleeding) for 123 blue marlin (Makaira nigricans ), 272 white marlin (Kajikia albida), and 132 sailfish (Istiophorus platypterus ) caught on natural baits rigged with one of the two hook types. In addition, we used pop-up satellite archival tags (PSATs) to follow the fate of 61 blue marlin caught on natural baits rigged with circle hooks or on a combination of artificial lure and natural bait rigged with J hooks. The frequencies of inter- nal hooking locations and bleeding were significantly lower with circle hooks than with J hooks for each of the three species and were signifi- cantly reduced for blue marlin caught on J hooks than for white marlin and sailfish taken on the same hook type. Analysis of the data received from 59 PSATs (two tags released prematurely) indicated no mortali- ties among the 29 blue marlin caught on circle hooks and two mortalities among the 30 blue marlin caught on J hooks (6.7%). Collectively, the hook location and PSAT data revealed that blue marlin, like white marlin and sailfish, derive substantial conserva- tion benefits from the use of circle hooks, and the negative impacts of J hooks are significantly reduced for blue marlin relative to the other two species. Manuscript submitted 6 April 2010. Manuscript accepted 22 July 2010. Fish. Bull. 108:433-441 (2010). The views and opinions expressed or implied in this article are those of the author (or authors) and do not necessarily reflect the position of the National Marine Fisheries Service, NOAA. Asymmetric conservation benefits of circle hooks in multispecies hillfish recreational fisheries: a synthesis of hook performance and analysis of blue marlin ( Makaira nigricans ) postrelease survival John E. Graves (contact author)1 Andrij Z. Horodysky2 Email address for contact author: graves@vims.edu 1 Department of Fisheries Science Virginia Institute of Marine Science College of William & Mary Rt. 1208 Greate Road Gloucester Point, Virginia 23062 2 Living Marine Resources Cooperative Science Center Department of Marine and Environmental Science Hampton University 100 E. Queen St. Hampton, Virginia 23668 Istiophorid billfishes in the Atlantic Ocean experience considerable fish- ing pressure and most stocks are overfished. The greatest source of fishing-induced mortality for istio- phorids results from the pelagic long- line fishery that targets tunas and swordfish; however, artisanal and rec- reational fisheries also represent sig- nificant sources of mortality for some species (Arocha and Ortiz, 2006). The United States National Marine Fisheries Service (NMFS) man- ages the recreational hillfish fishery with relatively large minimum sizes released to ensure that the major- ity of billfishes are released: 251 cm (99 in) lower jaw fork length (LJFL) for blue marlin (Makaira nigricans), 168 cm (66 in) LJFL for white marlin (Kajikia albida ), and 152 cm (60 in) LJFL for sailfish (Istiophorus platyp- terus). No recreational landings are allowed for longbill spearfish (Tet- rapturus pfluegeri), and no manage- ment measures currently exist for roundscale spearfish (T. georgii). As a result of these management mea- sures and changes in angler behavior promoting live release of these species (Ditton and Stoll, 2003), the U.S. rec- reational hillfish fisheries are primar- ily catch-and-release fisheries, and up to 99% of white marlin are released alive annually (Goodyear and Prince, 2003). However, not all billfishes that are released alive survive capture; postrelease mortality can be signifi- cant in some fisheries (Domeier et ah, 2003; Horodysky and Graves, 2005). A growing body of evidence indi- cates that the use of circle hooks can greatly reduce the incidences of in- ternal (deep) hooking, hook induced trauma (bleeding), and postrelease mortality of piscivorous fishes (Mu- oneke and Childress, 1994; Skomal et al., 2002; Cooke and Suski, 2004), including billfishes (see Serafy et ah, 2009). For istiophorid billfishes, live and dead natural baits rigged with J hooks reveal higher frequencies of in- ternal hooking locations and trauma for sailfish, striped marlin (K. audax), and white marlin than the same baits rigged with circle hooks (Prince et al., 2002, 2007; Domeier et al., 2003; Horodysky and Graves, 2005). Using pop-up satellite archival tags (PSATs) to follow the fate of released fish, Do- meier et al. (2003) noted a reduced but nonsignificant postrelease mortal- ity for striped marlin caught on live natural baits rigged on circle hooks, and Horodysky and Graves (2005) re- ported a highly significant reduction 434 Fishery Bulletin 108(4) in postrelease mortality of fish caught on circle hooks compared to those caught on J hooks. In response to the depleted stock condition of Atlantic billfishes and the re- duction in undesirable hooking locations and postrelease mortality resulting from the use of circle hooks, NMFS in 2008 implemented a management measure requiring the use of non-offset circle hooks in natural baits for all Atlantic billfish tournaments. During the rule-making process, NMFS received several comments stating that blue marlin have lower rates of deep hooking with J hooks than white marlin or sailfish, especially when caught on artificial lure and natural bait combinations rigged with J hooks, a common terminal tackle used in the Atlantic recreational fishery. It was also suggested that blue marlin released from the recreational fishery have high rates of survival. Little is known about the effects of hook type on postrelease survival of blue marlin. A preliminary study on the use of PSATs to investigate postrelease survival of blue marlin inferred survival for at least eight of nine individuals caught on lures or skirted baits with J hooks trolled at relatively high speeds in the Bermuda recreational fishery (Graves et ah, 2002). Flowever, there were no data to directly compare the postrelease mortality of blue marlin caught on natural baits or on a combination of artificial lure and natural bait rigged with either circle hooks or J-hooks. To gain insights into the relative conservation benefits of circle hooks in the recreational fishery for Atlantic blue marlin and other istiophorids, we compiled data on hooking location and the incidence of trauma (bleeding) for 123 blue marlin, 272 white marlin, and 132 sailfish caught on natural baits rigged with either J hooks or circle hooks. Fur- thermore, to estimate the postrelease mortality of blue marlin caught on natural baits with circle hooks or ar- tificial lure and natural bait combinations with J hooks, we deployed 61 PSATs to follow the fate of blue marlin caught on one of the two types of terminal tackle. Materials and methods Hooking location Information on hook type, hooking location, and trauma (the presence of bleeding) was collected by the authors for all blue marlin, white marlin, and sailfish caught on trolled plain (naked) ballyhoo ( Herniramphus brasil- iensis) baits during our PSAT tagging operations in the western North Atlantic from 2006 through 2009. Similar information was recorded by three cooperating charter captains (one from Oregon Inlet, NC, USA, and two from La Guaira, Venezuela) for billfishes caught during fishing operations in 2006. To accommodate captain and charter angler preferences, a variety of hook models and sizes were employed in the different fishing operations, but the most common J hook model was the Mustad 9175, size 7/0 (Mustad, Gjovik, Norway), and the most common circle hook model was the Eagle Claw L2004EWF, size 8/0 (Eagle Claw, Denver, CO). As is typical for this fish- ery, circle hooks were left exposed, rigged to the head of the bait, whereas J hooks were inserted through the mouth of the bait fish with the tip exiting the ventral surface (Fig. 1, A and B). Baits were trolled at approximately 6 nm/hr (11.1 km/ hr) during daylight hours on 30-50 lb (13.6-22.7 kg) class fishing tackle. As billfish approached the trolled bait, anglers would typically decrease tension on the line for periods of 4-10 s, “dropping back” the ballyhoo bait to the feeding billfish, which provided time for the Graves and Horodysky: Conservation benefits of circle hooks in multispecies recreational billfish fisheries 435 animal to ingest the bait before feeling tension on the line (see Jolley, 1974; Mather et al., 1975; Prince et al., 2007, for a description of this fishing method), although, in some instances the fish attacked trolled baits before an angler could reach the rod and drop back. The loca- tion of the hook and the presence or absence of bleeding (visible blood) was noted at the time of capture. Hooking locations were classified as external when all or part of the hook was visible outside of the fish’s mouth, includ- ing fish that were foul hooked (hooked in areas away from the fish’s mouth), and internal if no part of the hook was visible when the fish’s mouth was closed. Postrelease survival PSATs were attached to 61 blue marlin caught in rec- reational fisheries in the western Atlantic Ocean. Fish were caught by using J hooks and circle hooks that were rigged with natural baits consisting of ballyhoo or Span- ish mackerel (Scomberomorous maculatus) . Baits were rigged in a manner typical for the fishery. J hooks were inserted inside the baits, which were fished in combina- tion with a skirted artificial lure (e.g., Ilander [L & S Bait Company, Largo, FL], chugger [Mold Craft Lures, Pompano Beach, FL], or Seawitch [C & H Lures, Jack- sonville, FL]) attached directly ahead of the bait (Fig. 1, C and D). Circle hooks were rigged externally on the baits, either on top of the head or directly in front of the bait. In some cases, a small artificial lure (chugger) was placed between the circle hook and the bait. Blue marlin were caught on 30-130 lb (13.6-59.1 kg) class sportfishing tackle in waters off the United States mid-Atlantic coast; St. Thomas, U.S. Virgin Islands; Punta Cana, Dominican Republic; La Guaira, Venezuela; and Porto Seguro, Brazil, between September 2007 and October 2009 (Appendix 1). As is typical for the fishery, the vessel was maneuvered by the captain to assist the angler in the capture of the fish. Blue marlin were not brought alongside the vessel until they were considered to be sufficiently calm to allow accurate tag placement. The first 61 fish available to us were tagged with a Microwave Telemetry PTT 100 HR tag (Microwave Te- lemetry, Inc., Columbia, MD), programmed for release after 10 days. This tag model records temperature, pressure (depth), and light levels approximately every 90-120 s. The tags were rigged as described in Graves et al. (2002), and deployed as described in Horodysky and Graves (2005). The location of the hook and pres- ence or absence of bleeding was noted for each fish at the time of tagging. If practical, the hook was removed from the fish before its release. As is customary in this fishery, blue marlin that were unable to maintain their position upright in the water column were resuscitated by using the forward motion of the vessel to facilitate movement of water over the fish’s gills before release (Appendix 1). Survival of released blue marlin was inferred from temperature and depth profiles following the protocols of Horodysky and Graves (2005). Net displacement of each fish was calculated as the minimum straight line distance from the point of release to the point of tag pop-up. Cochran-Mantel-Haenszel (CMH) or Fisher’s exact tests were used to address the effect of the J hooks and circle hooks on hooking location, hook-in- duced trauma, and survival. A Yates correction for small sample size was applied in conducting CMH tests when expected cell values were less than 5 (Agresti, 1990). All statistical analyses were conducted in the Statistical Analysis System, vers. 9.1 (SAS Institute, Cary, NC). Bootstrapping simulations were performed to determine the 95% confidence intervals of the esti- mates of mortality after release by using the software developed by Goodyear (2002). Results Hooking location Hooking location and the presence or absence of bleeding were noted for 123 blue marlin, 272 white marlin, and 132 sailfish caught on natural baits rigged with either J hooks or circle hooks (Table 1). The incidence of internal hooking with J hooks ranged from 19.1% (blue marlin) to 44.4% (white marlin), and the frequency of internal hooking locations for fish caught on circle hooks was considerably lower, ranging from 1.8% (blue marlin) to 6.2% (sailfish). For blue marlin, white marlin, and sailfish, J hooks had a significantly higher probability of internal hooking locations than circle hooks (P<0.007, P<0.0001, P<0.001, respectively). The frequency of inter- nal hooking locations for J hooks in blue marlin (19.1%) was less than half the value observed for white marlin and sailfish (44.4% and 41.2%, respectively), and the difference between blue marlin versus white marlin and sailfish combined was significant (P<0.0014). The occurrence of trauma (bleeding) mirrored the pat- tern observed for internal hooking locations between the two hook types for each of the three billfishes. Across the three species, over 81% of the instances of bleeding (44/54) were associated with internal hooking loca- tions (7/9 blue marlin, 20/24 white marlin, and 16/17 sailfish). Bleeding of fish caught on circle hooks ranged from 0% in blue marlin to 2.5% in sailfish (Table 1). For blue marlin, white marlin, and sailfish, J hooks had a significantly higher probability of inducing bleeding than circle hooks (P<0.0141, P<0.001, P<0.0001, re- spectively). As with the occurrence of internal hooking locations with J hooks among the three species, the frequency of bleeding observed in blue marlin (13.2%) was less than half that observed in white marlin and sailfish (both 33.3%); blue marlin had significantly lower rates of bleeding resulting from the use of J hooks than white marlin and sailfish combined (P<0.027). Postrelease survival Sixty-one blue marlin were caught on natural baits rigged with circle hooks (30) or on a combination of artificial lure and natural bait rigged with J hooks (31), 436 Fishery Bulletin 108(4) Table 1 Hooking location (frequency; 95% confidence interval) and presence or absence of trauma (bleeding, not bleeding) for observed recreational catches of blue marlin (Makaira nigricans, zi = 123), white marlin (Kajikia albida, n-212), and sailfish (Istiophorus platypterus, n = 132), caught on natural baits rigged with either J hooks or circle hooks in the western North Atlantic Ocean. Hook location Trauma Species Hook type Internal Externa] Bleeding Not bleeding Blue marlin Circle 1 (1.8%; 0-5.4%) 54 (98.2%; 90.0-99.9%) 0 (0%; 0-2.0%) 55 (100%; 93.5-100%) “J” 13(19.1%; 10.6-28.5%) 55 (80.9%; 69.5-89.4%) 9 (13.2%; 6.2-23.6%) 59 (86.8%; 78.7-94.8%) White marlin Circle 4 (2.0%; 0. 6-5.0%) 196 (98.0%; 94.5-99.5%) 2 (1.0%; 0-2.4%) 198 (99.0%; 96.4-99.9%) “J” 32 (44.4%; 32.7-56.6 %) 40 (55.6%; 43.4-67.3%) 24 (33.3% ; 22.4-44.2%) 48 (66.7%; 54.6-77.3%) Sailfish Circle 5 (6.2%; 2.0-13.8%) 76 (93.8%; 86.0-97.8%) 2 (2.5%; 0.3-8. 6%) 79 (97.5%; 91.4-99.7%) “J” 21 (41.2%; 27.6-55.8%) 30 (58.8%; 44.3-72.4%) 17 (33.3%; 20.7-47.9%) 34(66.7%; 52.1-79.2%) tagged with Microwave Telemetry PSATs, and released. Three blue marlin were released off the U.S. mid-Atlan- tic coast; 26 off St. Thomas, U.S. Virgin Islands; 2 off Punta Cana, Dominican Republic; 21 off La Guaira, Venezuela, and 9 off Porto Seguro, Brazil (Appendix 1). Estimated fish weights ranged from 70 lb (31.8 kg) to 500 lb (227.3 kg) (mean=216.4 lb [98.2 kg]). Fight times (including tag attachment) ranged from 4 to 85 minutes (mean=19.4 minutes). After tag placement, eight fish exhibited difficulty maintaining an upright orientation alongside the boat and were resuscitated for periods ranging from one to ten minutes before their release (Appendix 1). All 61 PSATs reported after detachment from the fish. Two tags detached prematurely — both on the first day of deployment — and were excluded from survival analyses. The 59 tags that remained attached for the ten-day deployment period successfully transmitted between 18% and 96% (mean = 81%) of the archived data. Most of these PSATs remained at sea during the data transmission period which typically lasts about 30 days before the battery power is exhausted. How- ever, six tags washed ashore during the period of data transmission, resulting in reduced data reception from these tags. Two tags were recovered by beachcombers and returned to us, allowing recovery of 100% of the archived data. We inferred the survival of 57 of 59 fish (96.6%) that carried the tags for the 10-day deployment from analyses of pressure (depth) and temperature profiles over the 10-day tagging period (Appendix 1). Surviv- ing blue marlin exhibited multiple daily vertical move- ments as evidenced by the temperature and pressure (depth) profiles throughout the course of the ten-day tagging period. Many animals demonstrated a distinct diurnal patterning to their dives, remaining near the surface at night and making deep dives during the day (Fig. 2A). Net displacement ranged from 10 to 943 km (mean=226.9 km). The two mortalities inferred from the PSAT data occurred among the 30 blue marlin caught on artifi- cial lure and natural bait combinations rigged with J-hooks. Both individuals were hooked internally and bled profusely from the gill area at the time of capture. The archived data indicated that these in- dividuals sank to the bottom shortly after release and remained there for 48-96 hr, after which time a release mechanism was activated by the constant depth data in each PSAT and initiated tag release and data transmission (Fig. 2B). The two mortalities of blue marlin caught on artificial lure and natural bait combinations with J hooks resulted in an esti- mated postrelease mortality rate of 6.7%. The results of 10,000 bootstrap simulations at an underlying true mortality of 6.7% indicated that the approximate 95% confidence intervals (Cl) for the mortality estimates of blue marlin caught on J hooks for an experiment with 30 tags would range from 0% to 22% (with the methods of Goodyear, 2002). None of the 29 blue mar- lin caught on natural baits with circle hooks died during the ten day period, resulting in a postrelease mortality estimate of 0%, with corresponding 95% Cl ranges from 0% to 12.5%. The difference between the estimates of postrelease mortality for blue mar- lin caught on J-hooks and circle hooks was not sta- tistically significant (Fisher’s exact test: P=0.26). Discussion The goal of this study was to determine whether blue marlin derive similar conservation benefits from the use of circle hooks in the recreational fishery as has been previously reported for other istiophorid billfishes. A direct comparison of hooking locations with the use of J hooks and circle hooks in natural baits rigged as they are typically fished in the recreational fishery indicates that the use of circle hooks results in sig- Graves and Horodysky: Conservation benefits of circle hooks in multispecies recreational billfish fisheries 437 A Q. Q B Figure 2 Depth plots derived from pop-up satellite archival tag pressure data for two blue marlin ( Makaira nigricans ) caught in the western Atlantic recreational fishery. (A) Blue marlin no. 49 was caught on a naked ballyhoo bait with a circle hook. This fish survived for the ten day tagging period and exhibited a strong diel pattern, diving during the day and remaining in surface waters at night. (B) Blue marlin no. 41 was caught on an artificial lure (Ilander) and ballyhoo combination bait with a J hook and was bleeding profusely from the gills at the time of capture. This fish died and sank to the bottom shortly after release. The release mechanism on the tag was activated by constant depth measurements for 48 hours and the tag floated to the surface and began transmitting data. nificantly lower rates of internal hooking locations for blue marlin, white marlin, and sailfish. We observed incidences of internal hooking locations with circle hooks ranging from 1.8% (blue marlin) to 6.2% (sailfish). The value of 2.0% (;i=200) observed for white marlin is comparable to the value of 1.7% (n = 59) reported for white marlin caught on natural baits with circle hooks (Graves and Horodysky, 2008). The incidence of internal hook- ing locations observed for circle hooks in sailfish (6.2%) is slightly higher than that reported by Prince et al. (2002) for Pacific sailfish caught on trolled natu- ral baits with circle hooks (1.7%) but is within the range reported for circle hooks in live baits for both Atlantic sailfish (6-16%; Prince et al., 2007) and striped marlin (5-7%; Domeier et al., 2003). The use of J hooks in natural baits resulted in incidences of internal hook- ing ranging from 19.1% (blue marlin) to 44.4% (white marlin). The frequency of internal hooking in white marlin caught with natural baits rigged with J hooks (44.4%) is similar to the value report- ed for white marlin caught on J hooks (50%) by Horodysky and Graves (2005). Internal hooking locations for sailfish caught on J hooks rigged with natural baits (41.2%) are comparable to results for Pacific sailfish caught on trolled dead baits rigged with J hooks (46.8%), and fall within the range reported for Atlan- tic sailfish caught on live baits rigged with J hooks using a variety of dropback times (23-57%; Prince et al., 2007), as well as striped marlin caught on live baits with J hooks (28%; Domeier et al., 2003). The use of J hooks resulted in a tenfold increase in internal hooking locations relative to circle hooks for blue marlin and a twentyfold increase for white marlin and sailfish — a trend also noted in previous studies of istiophorid billfish (Prince et al., 2002, 2007; Domeier et al., 2003; Horodysky and Graves, 2005). Although the frequency of internal hooking locations was significantly higher for blue marlin, white marlin, and sailfish caught on J hooks than on circle hooks, the rate of internal hooking locations for blue marlin caught on J hooks was less than half of the values observed for white marlin and sailfish. In a study of postrelease mortality in the recreational blue marlin fishery off Bermuda, Graves et al. (2002) reported no internal hooking locations for the nine blue marlin caught on artificial lures or artificial lure and natural bait com- binations rigged with J hooks. The lower incidence of internal hooking locations for blue marlin caught on natural baits rigged with J hooks than for white mar- lin and sailfish caught on similar terminal tackle may result from interspecific differences in feeding ecology. Many billfishes follow trolled baits for a short time before striking, giving alert anglers an opportunity to pick up the rod and drop the bait back to the fish as it attacks. Dropback times of 5-10 s are common in the white marlin fishery (Mather et al., 1975; Jesien et al., 2006), and can be considerably longer in the sailfish live bait fisheries (Prince et al., 2007). By contrast, blue marlin are typically more aggressive feeders, often attacking the bait before anglers have an opportunity to react. When dropbacks are possible for this species, they are often of shorter duration than those for white marlin and sailfish, allowing less time for the bait to 438 Fishery Bulletin 108(4) be swallowed and hence for the hook to lodge in an undesirable location (Prince et al., 2007). In previous studies, internal or deep hooking loca- tions have been associated with an increased incidence of trauma in istiophorids and other large pelagic fishes (Domeier et al. 2003; Horodysky and Graves, 2005; Prince et al., 2007; Skomal, 2007). In the present study, trauma, as evidenced by bleeding, was significantly lower for blue marlin, white marlin, and sailfish caught on circle hooks than for those caught on J hooks. The low incidence of bleeding associated with natural baits rigged with circle hooks for white marlin that we ob- served (1.0%) concurs with the value of 1.7% reported for white marlin by Graves and Horodysky (2008) and the range reported for the congeneric striped marlin caught on live baits with circle hooks (3-4%, Domeier et al., 2003). The frequency of bleeding with circle hooks observed for sailfish (2.5%) in this study is slightly low- er than the values reported for Pacific sailfish caught on natural baits (6%) and for Atlantic sailfish caught on live baits with circle hooks (5-13%; Prince et al., 2002, 2007). We noted bleeding in 33% of the white marlin and 33% of the sailfish caught on natural baits with J hooks. This value is somewhat lower than those report- ed for white marlin (45%) and Pacific sailfish (56.8%) caught on trolled natural baits, and higher than those reported for striped marlin (21%) and Atlantic sailfish (21-25%) caught on live baits (Prince et al., 2002, 2007; Domeier et al., 2003; Horodysky and Graves, 2005). The incidence of bleeding in white marlin and sailfish caught on J hooks was significantly higher than that for blue marlin caught on the same terminal tackle and is consistent with an increased incidence of internal hooking rates in the former species observed in this and previous studies (Prince et al., 2002; 2007; Domeier et al., 2003; Horodysky and Graves, 2005). To directly follow the fate of blue marlin caught on trolled natural baits or on bait and lure combinations rigged with either circle or J hooks and released, we deployed 61 PSATs, of which 59 remained attached for the ten-day tracking period. All 29 blue marlin caught on circle hooks rigged with natural baits or bait and lure combinations survived for ten days after release. In a study of 59 white marlin caught on natural baits rigged with circle hooks, Graves and Horodysky (2008) reported a single mortality. Together, these studies reveal a very low level of postrelease mortality for bill- fishes caught on circle hooks in natural bait trolling fisheries. In contrast, Domeier et al. (2003) reported an adjusted rate of postrelease mortality of 17.4% for striped marlin caught from stationary vessels on live baits rigged with circle hooks. These results support the contention of Cooke and Suski (2004) that the magni- tude of the conservation benefits of circle hooks varies among species, gear types, and fisheries. PSAT depth records indicated that two of 30 blue marlin caught on artificial lure and natural combina- tions rigged with J hooks died after release, resulting in an estimated postrelease mortality of 6.7%. This value is approximately one-fifth of the postrelease mortality reported by Horodysky and Graves (2005) for 20 white marlin caught on natural baits rigged with J hooks (35%) and less than one-fourth of the value of 29.4% reported by Domeier et al. (2003) for 24 striped marlin caught on live baits rigged with J hooks. The reduction in postrelease mortality of blue marlin caught on natu- ral baits with J hooks relative to postrelease mortality of white marlin caught on the same terminal tackle parallels the reductions observed in the frequency of internal hooking locations and bleeding between these species. In this study, the use of circle hooks in natural baits resulted in significantly reduced internal hooking and bleeding for blue marlin, white marlin, and sailfish. Because blue marlin caught on J hooks experienced significantly lower incidences of internal hooking loca- tions and trauma than white marlin and sailfish caught on the same terminal tackle, the conservation benefit resulting from the use of circle hooks for blue marlin is less than that experienced by the other two species. This asymmetry was also evident in the analysis of postrelease survival. There was a trend for decreased postrelease mortality of blue marlin caught on natu- ral baits with circle hooks (0%) than for blue marlin caught on artificial lures and natural baits with J hooks (6.7%), but the difference was much smaller than that previously reported for white marlin caught on natural baits rigged with the two hook types (0% and 35%, respectively; Horodysky and Graves, 2005). Although these results provide support for recent management measures implemented by NMFS that require the use of non-offset circle hooks in natural baits for all Atlantic billfish tournaments, it is important to realize that the conservation benefits of this measure vary asymmetri- cally among the different billfish species. Acknowledgments We thank Captains J. Grant, R. Cox and J. Ross for compiling information on hook location and fish condi- tion. B. DeGabrielle, J. Tierney, J. Thiel, K. Neill, D. Dutton, J. Schratweiser, and many others provided assis- tance in catching and tagging blue marlin. M. Domeier and the International Game Fish Association provided logistical support. Funding for this study was provided by the Gulf States Marine Fisheries Commission (Bill- fish-2005-11), the National Marine Fisheries Service (NA07NMF4720295), and the Offield Family Foundation. This article is VIMS contribution number 3107. Literature cited Argesti, A. 1990. Categorical data analysis, 558 p. Wiley, New York. Arocha, F., and M. Ortiz. 2006. Standardized catch rates for blue marlin ( Mak - aira nigricans) and white marlin (Tetrapturus albidus ) Graves and Horodysky: Conservation benefits of circle hooks in multispecies recreational billfish fisheries 439 from the Venezuelan pelagic longline fishery off the Caribbean Sea and the western central Atlantic: period 1991-2004. ICCAT Coll. Vol. Sci. Pap. 59:315-322. Cooke, S. J., and C. D. Suski. 2004. Are circle hooks an effective tool for conserving marine and freshwater recreational catch and release fisheries? Aquat. Conservat. Mar. Freshw. Ecosyst. 14:299-326. Ditton, R. B., and J. R. Stoll. 2003. Social and economic perspective on recreational billfish fisheries. Mar. Freshw. Res. 54:545-554. Domeier, M. L, H. Dewar, and N. Nasby-Lucas. 2003. Mortality rate of striped marlin (Tetrapturus audax) caught with recreational tackle. Mar. Freshw. Res. 54:435-445. Goodyear, C. P. 2002. Factors affecting robust estimates of the catch- and-release mortality using pop-off tag technology. In Catch and release in marine recreational fisheries (J. A. Lucy and A. L. Studholme, eds.), p. 172-179. Am. Fish. Soc. Symp. 30, Bethesda, MD. Goodyear, C. P., and E. D. Prince. 2003. U.S. recreational harvest of white marlin. ICCAT Coll. Vol. Sci. Pap. 55:624-632. Graves, J. E., B. E. Luckhurst, and E. D. Prince. 2002. An evaluation of pop-up satellite tags for estimating postrelease survival of blue marlin ( Makaira nigricans) from a recreational fishery. Fish. Bull. 100:134-142. Graves, J. E., and A. Z. Horodysky 2008. Does hook choice matter? The effects of three circle hook models on post-release survival of white marlin. N. Am. J. Fish. Manag. 28:471-480. Horodysky, A. Z., and J. E. Graves. 2005. Application of pop-up satellite archival tag tech- nology to estimate postrelease survival of white marlin (Tetrapturus albidus ) caught on circle and straight-shank (“J”) hooks in the western North Atlantic recreational fishery. Fish. Bull. 103:84-96. Jesien, R. V., A. M. Barse, S. Smyth, E. D. Prince, and J. E. Serafy. 2006. Characterization of the white marlin ( Tetrapterus albidus) recreational fishery off Maryland and New Jersey. Bull. Mar. Sci. 79:647-657. Jolley, J. W., Jr. 1974. On the biology of Florida East Coast Atlantic sail- fish (Istiophorus platypterus). In Proceedings of the international billfish symposium, part 2: review and contributed papers (R. S. Shomura and F. Williams, eds.); August 1972, Kailua-Kona, HI, p. 81-88. NOAA Tech Rep. NMFS SSRF-675. Mather, F. J., Ill, H. L. Clark, and J. M. Mason Jr. 1975. Synopsis of the biology of the white marlin Tet- rapturus albidus Poey (1861). In Proceedings of the international billfish symposium, part 3: species syn- opses (R. S. Shomura and F. Williams, eds.); August 1972, Kailua-Kona, HI, p. 55-94. NOAA Tech Rep. NMFS SSRF-675. Muoneke, M. I., and W. M. Childress. 1994. Hooking mortality: a review for recreational fisheries. Rev. Fish. Sci. 2:123-156. Prince, E. D., M. Ortiz, and A. Venizelos. 2002. A comparison of circle hook and “J” hook per- formance in recreational catch-and-release fisheries for billfish. In Catch and release in marine recre- ational fisheries (J. A. Lucy and A. Studholme, eds.), p. 66-79. Am. Fish. Soc. Symp. 30, Bethesda, MD. Prince, E. D., D. Snodgrass, E. Orbesen, J. P. Hoolihan, J. E. Serafy, and J. E. Schratweiser. 2007. Circle hooks, ‘J’ hooks and drop-back time: a hook performance study of the south Florida recreational live- bait fishery for sailfish, Istiophorus platypterus. Fish. Manag. Ecol. 14:173-182. Serafy, J. E., D. W. Kerstetter, and P. H. Rice. 2009. Can circle hook use benefit billfishes? Fish Fish. 10:132-142. Skomal, G. B. 2007. Evaluating the physiological and physical conse- quences of capture on post-release survivorship in large pelagic fishes. Fish. Manag. Ecol. 14:81-89. Skomal, G. B., B. C. Chase, and E. D. Prince. 2002. A comparison of circle and straight hook perfor- mance in recreational fisheries for juvenile Atlantic bluefin tuna. In Catch and release in marine recre- ational fisheries (J. A. Lucy and A. Studholme, eds.), p. 57-65. Am. Fish. Soc. Symp. 30, Bethesda, MD. 440 Fishery Bulletin 108(4) Appendix 1 Summary information for 61 blue marlin ( Makaira nigricans) caught on different types of terminal tackle in the western Atlantic recreational fishery, tagged with pop-up satellite archival tags, and released. Fish were caught off Venezuela (VZ), the Domini- can Republic (DR), Virginia (VA), North Carolina (NC), the U.S. Virgin Islands (VI), and Brazil (BR) on natural baits or combina- tion baits consisting of an artificial lure (chugger [Chug], Ilander [I], or Sea Witch [Swi]), and natural bait rigged with J hooks ( J) or circle hooks (C). Hook locations were recorded as internal (I) or external (E). Fight time included the tagging process and, where applicable, resuscitation (Rduratlon m minutes), ancj trauma (bleeding) at the time of release was noted as present (Y) or absent (N). Also included are the percentage of archived data recovered, mean straight line displacement (MSLD) over the 10-day tag- ging period, and fate of the blue marlin (coded as live [L], dead [D], or premature release [-]). Fish Date Loca- tion Gear (lb) Established weight Hook type Hook location Fight time Trauma Data (%) MSLD (km) Fate 1 9/7/07 VZ 30 200 Chug/J I 55 N 82 172 L 2 3/17/08 VZ 50 175 I/J E 28 R3 N 81 139 L 3 3/28/08 DR 50 125 I/J E 15 R2 N 68 49 L 4 3/29/08 DR 50 100 C E 30 N 85 86 L 5 5/15/08 VZ 30 165 C E 18 N 84 30 L 6 5/15/08 VZ 30 100 Chug/C E 5 N 85 10 L 7 5/16/08 VZ 50 160 I/J E 5 Y 73 123 L 8 5/17/08 VZ 30 75 C E 15 N 85 87 L 9 6/15/08 VA 50 400 I/J E 56 N 85 546 L 10 6/22/08 NC 80 350 I/J E 35 N 88 943 L 11 8/8/08 VI 50 450 I/J E 18 N 90 502 L 12 8/10/8 VI 50 275 I/J I 5 N 90 646 L 13 8/10/08 VI 50 350 Chug/C E 8 N 92 349 L 14 8/11/08 VI 50 175 Chug/J E 11 N 83 107 L 15 8/11/08 VI 50 175 Chug/C E 4 N 88 343 L 16 8/12/08 VI 50 200 I/J I 10 Y 93 488 L 17 8/12/08 VI 50 125 I/J E 12 Y 87 328 L 18 8/17/08 NC 80 375 Swi/J E 40 N 90 709 L 19 9/7/08 VI 50 160 I/J E 20 N 91 191 L 20 9/7/08 VI 50 250 Chug/C E 15 N 68 20 L 21 9/7/08 VI 50 235 C E 22 N 92 319 L 22 9/8/08 VI 50 275 Chug/J I 55 N 90 120 L 23 9/9/08 VI 50 125 I/J I 14 N 93 238 L 24 9/10/08 VI 50 90 Chug/J I 20 R10 Y 72 68 L 25 9/10/08 VI 50 90 Chug/J E 13 N 88 264 L 26 9/11/08 VI 80 325 I/J E 35 N 100 197 L 27 9/15/08 VI 50 100 Chug/C E 12 N 92 234 L 28 9/25/08 VZ 30 70 C E 5 N 90 13 L 29 11/1/08 BR 130 300 Chug/C E 11 R2 N 85 119 L 30 11/2/08 VZ 30 150 Chug/C E 21 N 84 336 L 31 11/3/08 VZ 30 290 C E 85 N 94 171 L 32 11/4/08 BR 50 125 C E 5 N 74 284 L 33 11/15/08 VZ 30 250 C E 20 N 18 263 L 34 1 1/17/08 VZ 30 150 C E 20 N 87 135 L 35 11/22/08 VZ 30 150 C E 21 N 79 115 L 36 11/30/08 VZ 30 120 Chug/C E 36 R4 N 86 83 L 37 12/2/08 BR 130 200 Chug/J I 12 Y 93 47 D 38 12/2/08 BR 130 175 Chug/C E 5 N 36 289 L 39 12/9/08 VZ 30 125 Chug/C I 21 N 81 226 L 40 12/9/08 VZ 50 100 I/J E 6 N 90 118 L 41 12/9/08 VZ 50 200 I/J I 9 Y 61 109 D continued Graves and Horodysky: Conservation benefits of circle hooks in multispecies recreational billfish fisheries 441 Appendix 1 (continued) Fish Date Loca- tion Gear (lb) Established weight Hook type Hook location Fight time Trauma Data (%) MSLD (km) Fate 42 12/11/08 VZ 30 250 I/J E 18 N 83 107 L 43 12/11/08 BR 130 450 Chug/C E 10 N 85 218 L 44 12/12/08 BR 130 500 C E 15 N 90 233 L 45 12/12/08 BR 50 400 C E 12 N 91 216 L 46 1/2/09 BR 130 300 C E 12 N 88 19 L 47 1/2/09 BR 130 400 Chug/C E 10 N 94 329 L 48 2/7/09 VZ 80 145 Swi/J E 10 N 26 124 L 49 4/18/09 VZ 80 115 C E 14 N 85 96 L 50 5/6/09 VZ 50 265 I/J E 34 N 57 88 L 51 5/9/09 VZ 50 110 C E 10 N 100 90 L 52 7/6/09 VI 50 170 Chug/J E 25 R3 N 85 70 — 53 7/6/9 VI 50 300 Chug/J E 45 R1 N 80 196 L 54 7/9/09 VI 50 200 Chug/J E 15 N 86 143 L 55 8/1/09 VI 50 250 Chug/C E 15 N 38 109 L 56 8/8/09 VI 50 250 Chug/J E 15 N 31 379 L 57 8/30/09 VI 50 175 C E 20 N 93 80 — 58 9/2/09 VI 50 150 C E 18 N 94 214 L 59 9/4/09 VI 50 350 C E 16 N 96 463 L 60 9/30/09 VI 50 175 C E 14 N 90 271 L 61 10/7/09 VI 50 200 I/J E 40 N 92 385 L 442 Abstract — Settled juvenile blue rock- fish ( Sebastes mystinus) were collected from two kelp beds approximately 335 km apart off Mendocino in northern California and Monterey in central California. A total of 112 rockfish were collected from both sites over 5 years (1993, 1994, 2001, 2002, and 2003). Total age, settlement date, age at settlement, and birth date were determined from otolith microstructure. Fish off Mendocino settled mostly in June and fish off Monterey settled mostly in May (aver- age difference in settlement=23 days). Although the difference in the timing of settlement followed this same pat- tern for both areas over the five years, settlement occurred later in 2002 and 2003 than in the prior years of sam- pling. The difference in the timing of settlement was due primarily to differences in birth dates for the two areas. The time of settlement was positively related to upwelling and negatively related to sea level anomaly for most of the months before settlement. Knowledge of the timing of settlement has implications for design and placement of marine protected areas because protection of nursery grounds is frequently a major objective of these protected areas. The timing of settlement is also an important consideration in the planning of surveys of early recruits because mistimed surveys (caused by latitudinal differences in the timing of settlement) could produce biased estimates. Manuscript submitted 4 May 2010. Manuscript accepted 27 July 2010. Fish. Bull. 108:442-449 (2010). The views and opinions expressed or implied in this article are those of the author (or authors) and do not necessarily reflect the position of the National Marine Fisheries Service, NOAA. Influence of ocean conditions on the timing of early life history events for blue rockfish (Sebastes mystinus) off California Thomas E. Laidig Email address: tom.laidig@noaa.gov Fisheries Ecology Division Southwest Fisheries Science Center National Marine Fisheries Service National Oceanic and Atmospheric Administration 110 Shaffer Road Santa Cruz, California 95060 Successful recruitment of fishes into adult populations can have a major influence on population biomass. Within a species, recruitment events can vary widely in magnitude both temporally and spatially (Doherty and Fowler, 1994; Ralston and Howard, 1995; Caley et al., 1996). Under- standing this variability in recruit- ment strength is a critical goal in predicting adult population structure (Doherty and Fowler, 1994; Hidalgo et al., 2009). Knowledge of the spa- tial and temporal variability in recruitment of a population can lead to enhanced management of the spe- cies (e.g., to protection of important nursery areas) (Grorud-Colvert and Sponaugle, 2009). Variation in recruitment strength is influenced by multiple biotic and abiotic factors, including the mag- nitude of spawning biomass, preda- tion, available habitat, available prey, competition, temperature, upwelling, turbulence, water quality, and ocean currents (Peterman and Bradford, 1987; Ainley et al., 1993; Ralston and Howard, 1995; Hobson et al., 2001; Johnson et al., 2001; Sale et al., 2005; Hidalgo et al., 2009). Re- cruitment variability can occur from large-scale (e.g., El Nino or La Nina events, Pacific Decadal Oscillations) (Carr, 1991; Field and Ralston, 2005; Laidig et al., 2007) to small-scale pro- cesses (e.g., localized patchiness of suitable nursery habitats; fine-scale oceanographic events) (Sale et al., 2005; Johnson, 2006). Determining the relative influence of these pro- cesses on fish recruitment has been a goal for many studies (Peterman and Bradford, 1987; Yoklavich et al., 1996; Laidig et al., 2007). Annual recruitment of rockfishes ( Sebastes spp.) in the northeast Pa- cific can vary by orders of magnitudes between years (Ralston and Howard, 1995; Laidig et al., 2007). In addition to temporal variability in recruitment, latitudinal or spatial variability also has been observed for several rockfish species along the west coast (Sakuma et al., 2006). Variable recruitment leads to greater uncertainty in the prediction of year-class strength. Rockfishes give birth to larvae (par- turition) that survive in the plankton for several months before settling into nursery or adult habitats (Carr, 1991; Love et al., 2002; Ammann, 2004). For several rockfish species, it has been demonstrated that the strength of the year class is determined during this planktonic stage (Ralston and Howard, 1995; Laidig et al., 2007; Wilson et al., 2008) and that it can be affected further by postsettlement mortality from predation (Hobson et al., 2001). In this study, the recruitment of ju- venile blue rockfish (S. mystinus) from the pelagic environment to nearshore kelp beds was examined in two geo- graphic regions along the California coast. Blue rockfish are an important commercial and recreational species in California ranging from at least British Columbia to northern Baja California (Love et al., 2002). Otolith microstructure was used to estimate Laidig: Influence of ocean conditions on the timing of early life events for Sebastes mystmus 443 Table 1 Location, year, sampling date, number of samples col- lected, and size range of blue rockfish (Sebastes mystinus) examined for settlement marks in otoliths. Location Year Sample date Number of samples Size range (mm standard length) Mendocino 1993 2 Jun 11 37-48 Monterey 1993 3 Jun 11 38-48 Mendocino 1994 26 Jul 13 44-52 Monterey 1994 20 Jul 13 42-60 Mendocino 2001 11 Sep 10 48-70 Monterey 2001 5 Sep 10 53-68 Mendocino 2002 13 Aug 12 47-66 Monterey 2002 18 Jul 13 45-58 Mendocino 2003 4 Sep 10 58-73 Monterey 2003 1 Aug 9 47-59 the timing of settlement, birth dates, and length of the pelagic larval and juvenile stages of blue rockfish. Upwelling indices and sea level anomalies were used to evaluate the influence of oceanographic conditions on the timing of recruitment (settlement) for blue rockfish in the two regions. Materials and methods Juvenile blue rockfish were collected in two areas, approximately 335 km apart — one off northern Califor- nia in Mendocino County (39°14'N lat., 123°46'W long.) and the other off central California along the southern edge of Monterey Bay in Monterey County (36°38'N lat., 121°55'W long.; Fig. 1). Each area is typified by large kelp beds that extend from shore to approximately 20 m depth. The Mendocino site is dominated by bull kelp ( Nereocystis luetkeana) and the Monterey site by giant kelp ( Macrocystis pyrifera). These kelp-bed areas com- prise high-relief bedrock interspersed with low-relief cobble and sand areas. Both beds are exposed to open ocean conditions, although the Monterey site is slightly buffered from southerly seas by the tip of the Monterey Peninsula. Fish were collected throughout the kelp beds by div- ers using small spears at depths of 5-20 m and were frozen for later analysis. Fish were collected during late spring and summer at both sites during five noncon- secutive years (1993, 1994, 2001-2003). Years for data analysis were selected on the basis of three criteria: 1) there were at least nine individuals collected from each area; 2) sampling dates at each study site in a particular year were reasonably close (approximately one month or less apart); and 3) samples were collected after 1 June to allow for complete settlement of the juveniles (Table 1). 39°N i Mendocino Pt. Arena V 38°N - \ San Francisco • Pacific Ocean 37°N XX 50 .km Monterey -124°N -1 23°N -122°N Figure 1 Map of the two study areas along the California coast. Black stars indicate areas where samples of blue rockfish (Sebastes mystinus) were collected for otolith analysis to determine the timing of early life history events. Otolith data Sagittal otoliths were removed and ages were deter- mined visually by counting growth increments with a compound microscope at 1000 x magnification (Laidig et al., 1991), beginning at the first increment after the extrusion check (a mark in the otolith formed when the larvae were released from their mother). No validation of the rate of deposition of these growth increments was performed during this study, and none was available from the literature. However, based on validation studies for other co-occurring rockfish species, such as shortbelly rockfish (S. jordani [Laidig et al., 1991]), bocaccio (S. paucispinis) , chilipepper ( S . goodei ), widow rockfish (S. entomelas), and yellowtail rockfish (S. flavidus [Wood- bury and Ralston, 1991]), increments were assumed to be deposited daily. Also, the calculated birth dates of the blue rockfish in this study occurred within the months of larval release for this species (Wyllie-Echeverria, 1987). The duration of the pelagic larval and juvenile stages, age at settlement (the total duration of pelagic larval and juvenile stages), settlement date, and birth date were calculated by identifying specific marks in the otolith that indicate transitions from one life stage to 444 Fishery Bulletin 108(4) Postsettleinent increments Pelagic increments Settlement mark Figure 2 Growth increments in the otolith of a blue rockfish ( Sebastes mystinus). The settlement mark and increments produced before (pelagic) and after (post) settlement are indicated. the next. Transformation from the larval to the juvenile stage was ascertained by the oc- currence of secondary growth primordia (areas of new increment growth that form away from the otolith core; Laidig et ah, 1991). This trans- formation occurs during the planktonic stage before settlement. Duration of the pelagic juve- nile stage was estimated as the number of in- crements occurring from the secondary growth primordia to the settlement mark. Settlement date was calculated by subtracting the number of increments formed after the settlement mark from the collection date, and birth date was determined by subtracting the total number of increments in the otolith from the collection date. The otolith increments deposited during the pelagic juvenile stage followed a regular pattern, increasing in width with fish age (Fig. 2). The settlement mark was the increment where a change in the depositional pattern of the increments was observed after the pelagic juvenile stage; typically the increment mark- ing settlement is narrower than the preceding increments (Amdur, 1991). To be considered the settlement mark, this change in increment width had to be visible from the settlement mark to the outer edge of the otolith and not just in one section. Often the settlement mark was evident in the otolith as a dark ring of numerous closely spaced increments. After the settle- ment mark, increment widths followed no consistent growth pattern. Oceanographic data Upwelling data were derived from monthly sea level pres- sure fields provided by the U.S. Navy Fleet Numerical Meteorology and Oceanography Center (data acquired from NMFS, Environmental Research Division, South- west Fisheries Science Center at http://www.pfeg. noaa.gov, accessed July 2009) measured off Mendocino (39°11'N lat., 123°58'W long.) and Monterey (36°47'N lat., 122°24,W long.), California. Sea level anomaly data (adjusted for local atmospheric conditions) were collected from shore stations at Humboldt Bay (40°46'N lat., 124°13'W long.) and Monterey (36°36'N lat., 121°53'W long.), California, and monthly means were obtained from the University of Hawaii Sea Level Center. These data represent a measure of change in sea level height over time and reflect water movement a positive anom- aly was associated with poleward flow and a negative anomaly was associated with equatorward flow. Two-way analysis of variance was used to test the hypotheses that mean settlement date, mean birth date, mean duration of pelagic larval and juvenile stages, and mean settlement age did not differ significantly between the two study areas. Principal components analysis (PCA) was used to evaluate the relationship among the otolith data (settlement date, birth date, and settlement age) and the oceanographic variables (monthly average upwelling and sea level anomalies for January-June) for both areas. Canonical correlation analysis (CCA) of the oceanographic and otolith data was used to ex- amine the possible causes of changes in interannual settlement dates. Resuits A total of 112 otoliths were examined (56 otoliths each from Mendocino and Monterey; Table 1). Sampling dates ranged from 2 June to 11 September each year, vary- ing from 1 to 34 days between the two study areas in a particular year. Fish sizes ranged from 37 to 73 mm standard length (SL) for Mendocino and from 38 to 68 mm SL for Monterey. Average birth dates varied annually, with a differ- ence of 7-30 days between locations and an average of 19 days difference over all years (Fig. 3). Birth dates of Mendocino rockfish were always later in the year than those of rockfish from Monterey (significantly different in 1994, 2001, and 2002; P< 0.05). Birth dates ranged from 4 January to 25 March for Mendocino fish (7 Feb- ruary average) and from 22 December to 9 March for Monterey fish (19 January average). Average settlement dates were significantly different (PcO.001) each year between the two locations, with fish from Mendocino always settling later in the year than fish from Monterey (23 days average; Fig. 4). The difference in settlement between the two sites varied from an average of 17 to 29 days. The earliest date of Laidig: Influence of ocean conditions on the timing of early life events for Sebastes mystinus 445 settlement was 17 May 1993 for fish off Mendocino and 23 April (occurring in both 2001 and 1993) for fish from Monterey; the latest settlement date was 1 July 2002 for fish off Mendocino and 4 June 2002 for fish off Monterey. Settlement dates were later in 2002 and 2003 and earlier in 1993 than in the other years. The duration of the larval and pe- lagic juvenile stages was not signifi- cantly different between the two areas in any particular year (Figs. 5 and 6). The average duration for the lar- val stage was 69 days (range = 41-100 d) for fish off Mendocino and 68 days (range = 47-90 d) for fish off Mon- terey. The average duration for the pelagic juvenile stage was 56 days (range=26-90 d) for fish off Mendoci- no and 52 days (range = 14-81 d) for fish off Monterey. Duration of pelagic larval and juvenile stages generally was longer for fish born in the 2000s than for fish born in the 1990s. Fish settled at younger ages at both study sites in the 1990s than in the 2000s (Fig. 7). Age at settlement was signifi- cantly different (P<0.05) between the two locations only in 2003; fish from Mendocino settled at an older age than those from Monterey. From the PCA, 56% of the variabili- ty in otolith data was explained by the first eigenvector for fish off Mendocino and Monterey. This vector was char- acterized by the inverse relationship of birth date and settlement age, i.e., the earlier the birth date, the older the settlement age. The second eigen- vector explained 41% and 39% of the variability in otolith data for fish off Mendocino and Monterey, respectively; this vector was characterized by later settlement dates and birth dates. The first eigenvector for oceanographic data explained 60% of the variability from Mendocino and 61% of the vari- ability in data from Monterey; this vector was associated with the inverse relationship of upwelling and sea level during the months of March through June off Mendocino and May and June off Monterey. The second eigenvector in the oceanographic data was related to the inverse relationship between upwelling and sea level in April at both sites and explained 30% and 28% of the variability in data from Men- docino and Monterey, respectively. ii March Q1 -j 19 February 09 20 January ♦ Mendocino ■ Monterey 1993 1994 # 2001 2002 2003 Year Figure 3 Average birth dates (in calendar days), back-calculated from otolith data, for blue rockfish ( Sebastes mystinus) from two study areas in California. Black stars indicate significant differences (P<0.05) between study areas for a particular year. Dashed lines represent the 5-year average for each site. Vertical bars represent one standard error. Horizontal lines on the y axis separate months. 29 i 19 ♦ Mendocino ■ Monterey June 09 30 20 May 10 * * ★ * * 1993 1994 2001 Year ★ * ★ * I 2002 2003 Figure 4 Average settlement dates (in calendar days), back-calculated from oto- lith data, for blue rockfish (Sebastes mystinus ) from two study areas in California. Black stars indicate significant differences (PcO.001). Dashed lines represent the 5-year average for each site. Vertical bars represent one standard error. Horizontal lines on the y axis separate months. 446 Fishery Bulletin 108(4) 78 76 ♦ Mendocino b Monterey CO > 03 C o 03 TJ "a3 > Q) CD 2 0 > < 74 72 70 68 66 64 62 - 60 -I t 1993 1994 ♦ ♦ a- Year 2001 2002 2003 Figure 5 Average larval duration, calculated from otolith data, for blue rockfish (Sebastes mystinus) from two study areas off California. Dashed lines represent the 5-year average for each site. Vertical bars represent one standard error. 75.0 70.0 65.0 60.0 55 0 50.0 45.0 40.0 35.0 30.0 ♦ Mendocino ■ Monterey 1993 1994 2001 Year 2002 2003 Figure 6 Average pelagic juvenile duration, calculated from otolith data, for blue rockfish ( Sebastes mystinus) from two study areas off California. Dashed lines represent the 5-year average for each site. Vertical bars represent one standard error. Results from the CCA were similar for data from both study areas (Fig. 8). For Mendocino, settlement dates later in the year (later than the average date for a particular year) were positively related to upwelling in February, May, and June, and inversely related to sea level anomalies in February, March, and May. For Monterey, later settlement dates were positively related to upwell- ing in all months, except April, and neg- atively related to sea level anomalies in all months, except April. There was an inverse relationship between birth date and settlement age for both sites, where- by an early birth date corresponded with an older settlement age. Discussion The timing of settlement of blue rockfish was related primarily to the timing of birth. Blue rockfish off Mendocino that settled on average three weeks later than fish off Monterey also were born, on aver- age, about three weeks later than fish off Monterey. The age at settlement did not differ significantly between sites, nor did the duration of the larval or pelagic juvenile stages. Pasten et al. (2003) also attributed differences in settling date to time of parturition because early settlers of Sebastes inermis to seagrass beds in Japan were born earlier than late-set- tling fish. These researchers concluded that there was an ideal size for settle- ment and for active migration. Sogard et al. (2008) studied the maternal effects of rockfishes on their progeny and postu- lated that factors influencing the time of parturition also influenced recruitment success. It must be noted that parturi- tion dates for blue rockfish in the present study were calculated from surviving juveniles. Those juveniles that did not survive may have been born at a differ- ent time and therefore would change the average parturition date (Woodbury and Ralston, 1991; Yoklavich et al., 1996). However, because no estimate of early larval or juvenile mortality exists for blue rockfish, I could not adjust for this effect. Time of spawning and parturition of many fish species vary by latitude. Parturition dates for numerous east- ern Pacific rockfish species occur later in northern study areas off Washington and Alaska than in areas off California Laidig: Influence of ocean conditions on the timing of early life events for Sebastes mystmus 447 145.0 i 140.0 - « 135.0 H E 130.0 - ® a> Z 125.0 - to (1) D) 120.0 - > 1 15.0 H < 110.0 H 105.0 ♦ Mendocino ■ Monterey 1993 1994 # 2001 2002 2003 Year Figure 7 Average age at settlement, calculated from otolith data, for blue rockfish ( Sebastes mystinus) from two study areas off California. Black stars indicate significant differences (P<0.05). Dashed lines represent the 5-year average for each site. Vertical bars represent one standard error. (Wyllie-Echeverria, 1987). Plaza et al. (2004) determined that parturi- tion of S. inermis occurred later in the season at the more northerly sites in the western Pacific and suggested that this difference may be related to environmental cues. Vinagre et al. (2008a, 2008b) reported a latitudi- nal gradient in time of spawning for European sea bass (Dicentrarchus labrax ) and common sole ( Solea solea) off Portugal and that spawning onset occurs earlier in the south. These re- searchers suggested that warm wa- ter temperatures in winter at lower latitudes or differences in photoperiod may influence the onset of spawning. The latitudinal difference in the timing of settlement and parturition in blue rockfish could be an adap- tation by blue rockfish to oceanic conditions. Bograd et al. (2009) cal- culated the spring transition index (i.e., the beginning of the upwelling season) and found that upwelling be- gan on average 20 days earlier off Monterey than off Mendocino. This difference is similar to the 23-day average difference in settlement timing between these two areas. It is possible that blue rockfish have adapted to this difference in the timing of upwelling. Interannual variability in parturition dates has been noted by researchers, but specific causes for this vari- ability are still unknown. Woodbury and Ralston (1991) observed variations in the timing of parturition for five species of rockfishes over six years. Early parturition dates have been related to increased maternal age and size for some rockfish species (Bobko and Berkeley, 2004; Plaza et al., 2004; Sogard et al., 2008). Plaza et al. (2004) proposed that the time of parturition could be influenced by effects of temperature on gestation times. Carr (1991) suggested that upwelling may influence the time of parturition. Interestingly, positive upwelling in January and February in this study was correlated with settlement date and this finding strengthens Carr’s speculation. Interannual variability in settlement date was re- lated to upwelling and sea level height in this study. Settlement occurred later in years when upwelling was stronger and sea level anomaly was negative (i.e., there was an equatorward flow of cold water). Stron- ger upwelling may have transported juvenile rockfish farther offshore, making their return to the nearshore more difficult and causing successful recruitment to occur later in the season (Ainley et al., 1993; Larson et al., 1994; Sakuma et al., 2006; Wilson et al., 2008). Years with reduced or negative upwelling should re- sult in onshore transport and thus earlier settlement to nearshore areas. Cold water produced by upwelling or flow from the north may result in slow growth of juvenile rockfishes (Boehlert and Yoklavich, 1983). These slow-growing individuals may migrate slowly to the nearshore environment. If prey are plentiful, the fish may remain in these food-rich waters longer or, if the quality of the prey items was low, the fish may exhibit reduced growth rates (Boldt and Rooper, 2009). Both of these factors could lead to later settle- ment of individuals. Large-scale oceanographic processes (covering 100s of kilometers) can be important in determin- ing recruitment strength over a broad area. Field and Ralston (2005) concluded that the synchrony in year-to-year recruitment for three rockfish species along the west coast of North America was caused by large-scale ocean processes (i.e., El Nino and Pa- cific Decadal Oscillation). Ralston and Howard (1995) reported similar recruitment strengths in blue and yellowtail rockfishes from two areas off California over a 10-year period and suggested large-scale ocean processes that affect sea surface temperatures (such as El Nino) are primarily responsible for the recruit- ment variation. Laidig et al. (2003) estimated simi- lar ages and growth curves for adult blue rockfish from Mendocino and Monterey, which indicated that the fish from these areas are influenced by similar oceanographic conditions. Understanding the causal relationship of biologic and oceanographic factors on recruitment dynam- 448 Fishery Bulletin 108(4) ics has implications for fisheries management. For instance, the design and placement of marine pro- tected areas can be augmented to include areas where juvenile rockfishes are recruited and grow, increas- ing the likelihood of sustained production for those rockfish species. Recruitment dynamics also need to be considered when conducting surveys of rockfish populations early in their life. These surveys can lead to enhanced recruitment predictions for the fishery. However, without information on the spatial and tem- poral occurrence of recruitment, surveys could be mistimed in a particular region and lead to biased estimates. Results of the present study indicate that coastwide recruitment surveys should be conducted from south to north to allow for later spawning times with increased latitude. Acknowledgments The work of all the divers who helped over the five years, especially T. Chess, T. Hobson, D. Howard, D. VenTresca, A. Ammann, and K. Baltz is greatly appreci- ated. M. Yoklavich and D. Watters offered helpful sug- gestions during this study. I also thank A. Ammann, S. Lonhart, K. Sakuma, B. Wells, M. Yoklavich, and three anonymous reviewers for their helpful comments of this manuscript. 1 0.75 0.5 0.25 0 -0.25 c -0.5 o t -0.75 8 -1.25 -1 -0.75 -0.5 -0.25 0 0.25 0.5 0 75 1 1.25 o § 1.25 T3 C O -I o 1 0) CO 0.75 0.5 0.25 0 -0.25 -0.5 -0.75 -1 -1 -0.75 -0.5 -0 25 0 0.25 0.5 0 75 1 1.25 First canonical correlation Figure 8 Comparison of first and second canonical correlation coefficients from oceano- graphic and otolith data at Monterey and Mendocino study sites. For each combined acronym, U=upwelling anomaly, L = sea level anomaly, and month is indicated. BD=birth date; SA=settlement age; SD = settlement date. Mendocino _ : BD _ UApril UJanuary UMarch UMay SD : LMaLr&?bruary ■ LMay LJanuary LJune — i 1 1 — i 1 r-~ i — LApril i 1 1 UFebruary UJune SA ~ i 1 1 i 1 1 Monterey - SA - UMarch LMay UApril LJune uFeW» SD LMarch . , LFebruary LJanuary LApril UJanuary i i i i i i i i i BD — i — i i — |— i i r i i Literature cited Ainley, D. G., W. J. Sydeman, R. H. Parrish, and R. H. Lenarz. 1993. Oceanic factors influenc- ing distribution of young rockfish ( Sebastes ) in central California: A predator’s perspective. CalCOFI Rep. 34:133-139. Amdur, J. 1991. The validation of a settlement mark on the otoliths of the blue rockfish Sebastes mystinus. M.S. thesis, 49 p. San Francisco State Univ., San Francisco, CA. Ammann, A. J. 2004. SMURFs: standard monitoring units for the recruitment of temper- ate reef fishes. J. Exp. Mar. Biol. Ecol. 299:135-154. Bobko, S. J., and S. A. Berkeley. 2004. Maturity, ovarian cycle, fecundity, and age-specific partu- rition of black rockfish, ( Sebastes melanops). Fish. Bull. 102:418-429. Boehlert, G. W., and M. M. Yoklavich. 1983. Effects of temperature, ration, and fish size on growth of juvenile black rockfish, Sebastes mela- nops. Environ. Biol. Fishes 8:17- 28. Bograd, S. J., I. Schroeder, N. Sarkar, X. Qiu, W. J. Sydeman, and F. B. Schwing. 2009. Phenology of coastal upwelling in the California Current. Geo- phys. Res. Let. 36: DOI: 10.1029/ 2008GL035933. Boldt, J. L., and C. N. Rooper. 2009. Abundance, condition, and diet of juvenile Pacific ocean perch ( Sebastes alutus ) in the Aleutian Islands. Fish. Bull. 107:278- 285. Caley, M. J., M. H. Carr, M. A. Hixon, T. P. Hughes, G. P. Jones, and B. A. Menge. 1996. Recruitment and the local dynamics of open marine pop- ulations. Ann. Rev. Ecol. Syst. 27: 477-500. Laidig: Influence of ocean conditions on the timing of early life events for Sebastes mystinus 449 Carr, M. H. 1991. Habitat selection and recruitment of an assemblage of temperate zone reef fishes. J. Exp. Mar. Biol. Ecol. 146:113-137. Doherty, P., and A. Fowler. 1994. Demographic consequences of variable recruitment to coral reef fish populations: A congeneric comparison of two damselfishes. Bull. Mar. Sci. 54:297-313. Field, J. C., and S. Ralston. 2005. Spatial variability in rockfish ( Sebastes spp. ) recruit- ment events in the California Current System. Can J. Fish Aquat. Sci. 62:2199-2210. Grorud-Colvert, K., and S. Sponaugle. 2009. Larval supply and juvenile recruitment of coral reef fishes to marine reserves and non-reserves of the upper Florida Keys, USA. Mar. Biol. 156:277-288. Hidalgo, M., J. Tomas, J. Moranta, and B. Morales-Nin. 2009. Intra-annual recruitment events of a shelf species around an island system in the NW Mediterranean. Es- tuar. Coast. Shelf Sci. 83:227-238. Hobson, E. S., J. R. Chess, and D. F. Howard. 2001. Interannual variation in predation of first- year Sebastes spp. by three northern Californian predators. Fish. Bull. 99:292-302. Johnson, D. W. 2006. Density dependence in marine fish populations revealed at small and large spatial scales. Ecology 87:319-325. Johnson, K. A., M. M. Yoklavich, and G. M. Cailliet. 2001. Recruitment of three species of juvenile rockfish ( Sebastes spp.) on soft benthic habitat in Monterey Bay, California. CalCOFI Rep. 42:153-166. Laidig, T. E., J. R. Chess, and D. F. Howard. 2007. Relationship between abundance of juvenile rock- fishes ( Sebastes spp.) and environmental variables documented off northern California and potential mech- anisms for the covariation. Fish. Bull. 105:39-48. Laidig, T. E., D. E. Pearson, and L. L. Sinclair. 2003. Age and growth of blue rockfish ( Sebastes mysti- nus) from central and northern California. Fish. Bull. 101:800-808. Laidig, T. E., S. Ralston, and J. R. Bence. 1991. Dynamics of growth in the early life history of shortbelly rockfish, Sebastes jordani. Fish. Bull. 89:611-621. Larson, R. J., W. H. Lenarz, and S. Ralston. 1994. The distribution of pelagic juvenile rockfish of the genus Sebastes in the upwelling region off central California. CalCOFI Rep. 35:175-221. Love, M. S., M. Yoklavich, and L. Thorsteinson. 2002. The rockfishes of the northeast Pacific, 405 p. Univ. California Press, Berkeley, CA. Pasten, G. P., S. Katayama, and M. Omori. 2003. Timing of parturition, planktonic duration, and settlement patterns of the black rockfish, Sebastes inermis. Environ. Biol. Fishes 68:229-239. Peterman, R. M., and M. J. Bradford. 1987. Wind speed and mortality rate of marine fish, the northern anchovy ( Engraulis mordax). Science 235:354-356. Plaza, G., S. Katayama, and M. Omori. 2004. Spatio-temporal patterns of parturition of the black rockfish Sebastes inermis in Sendai Bay, north- ern Japan. Fish. Sci. 70:256-263. Ralston, S., and D. F. Howard. 1995. On the development of year-class strength and cohort variability in two northern California rockfishes. Fish. Bull. 93:710-720. Sakuma, K. M., S. Ralston, and V. G. Wespestad. 2006. Interannual and spatial variation in the distribu- tion of young-of-the-year rockfish ( Sebastes spp.): expand- ing and coordinating a survey sampling frame. CalCOFI Rep. 47:127-139. Sale, P. F., B. S. Danilowicz, P. J. Doherty, and D. McB. Williams. 2005. The relation of microhabitat to variation in recruit- ment of young-of-year coral reef fishes. Bull. Mar. Sci. 76:123-142. Sogard, S. M., S. A. Berkeley, and R. Fisher. 2008. Maternal effects in rockfishes Sebastes spp.: a comparison among species. Mar. Ecol. Prog. Ser. 360:227-236. Vinagre, C., R. Amara, A. Maia, and H. N. Cabral. 2008a. Latitudinal comparison of spawning season and growth of 0-group sole, Solea solea ( L . ). Estuar. Coast. Shelf Sci. 78:521-528. Vinagre, C., T. Ferreira, L. Matos, M. J. Costa, and H. N. Cabral. 2008b. Latitudinal gradients in growth and spawning of sea bass, Dicentrarchus labrax, and their relation- ship with temperature and photoperiod. Estuar. Coast. Shelf Sci. 81:375-380. Wilson, J. R., B. R. Broitman, J. E. Caselle, D. E. Wendt. 2008. Recruitment of coastal fishes and oceanographic variability in central California. Estuar. Coast. Shelf Sci. 79:483-490. Woodbury, D., and S. Ralston. 1991. Interannual variation in growth rates and back- calculated birthdate distributions of pelagic juvenile rockfishes (Sebastes spp.) off the central California coast. Fish. Bull. 89:523-533. Wyllie-Echeverria, T. 1987. Thirty-four species of California rockfishes: Maturity and seasonality of reproduction. Fish. Bull. 85:229-250. Yoklavich, M. M., V. J. Loeb, M. Nishimoto, B. Daly. 1996. Nearshore assemblages of larval rockfishes and their physical environment off central California during an extended El Nino event, 1991-1993. Fish. Bull. 94:766-782. 450 Abundance and distribution of Atlantic sturgeon (Acipenser oxyrinchus ) within the Northwest Atlantic Ocean, determined from five fishery-independent surveys Email address for contact author: kdunton@notes.cc.sunysb.edu 1 School of Marine and Atmospheric Sciences Stony Brook University Stony Brook, New York 11794-5000 2 New York State Department of Environmental Conservation Division of Fish, Wildlife and Marine Resources Bureau of Marine Resources 205 North Belle Mead Road, Suite 1 East Setauket, New York 11733 Abstract — A lack of knowledge of how oceanic habitat is used by juvenile marine migrant Atlantic sturgeon (Acipenser oxyrinchus ) is hindering conservation measures directed at restoring severely depleted popula- tions. Identifying the spatial distribu- tion of Atlantic sturgeon is necessary to identify critical habitat and appro- priate management actions. We used five fishery-independent surveys to assess habitat use and movement of Atlantic sturgeon during their marine life stage. The size distribu- tion ranged from 56 to 269 cm total length (mean = 108 cm). Ninety-eight percent of all Atlantic sturgeon were smaller than 197 cm — a size that indi- cated the majority were immature. The pattern of habitat use revealed concentration areas and potential migration pathways used for north- erly summer and southerly winter migrations. Atlantic sturgeon were largely confined to water depths less than 20 m and aggregations tended to occur at the mouths of large bays (Chesapeake and Delaware bays) or estuaries (Hudson and Kennebec rivers) during the fall and spring and to disperse throughout the Mid- Atlantic Bight during the winter. In most surveys depth, temperature, and salinity were significantly related to the distribution of Atlantic sturgeon. Knowledge of their habitat and move- ments can be used to devise spatially based conservation plans to minimize bycatch and to enhance population recovery. Manuscript submitted 11 September 2009. Manuscript accepted 28 July 2010. Fish. Bull. 108:450-465 (2010). The views and opinions expressed or implied in this article are those of the author (or authors) and do not necessarily reflect the position of the National Marine Fisheries Service, NOAA. Keith J. Dunton (contact author)1 Adrian Jordaan1 Kim A. McKown2 David O. Conover1 Michael G. Frisk1 The Atlantic sturgeon ( Acipenser oxy- rinchus) is a long-lived anadromous fish with a historic range from Ham- ilton Inlet on the coast of Labrador to the Saint Johns River in Florida (Smith and Clugston, 1997). A major commercial fishery once existed throughout the historic range and estimated U.S. landings peaked at 3.3 million kg in 1890 (Smith and Clugston, 1997). Unable to support such intensive fishing, Atlantic stur- geon populations collapsed throughout the eastern seaboard by 1901 (Secor et al., 2002). During the late 1900s, there was a brief re-emergence of the Atlantic sturgeon fishery in New York and New Jersey (Kahnle et al., 2007) and landings peaked at 125,000 kg in the late 1980s (Waldman et al., 1996; Bain et al., 2000). In 1990 the Atlan- tic States Marine Fisheries Commis- sion (ASMFC) developed a fishery management plan for the conservation and restoration of Atlantic sturgeon in order to restore population levels that would support harvests at 10% of the historical peak landings (ASMFC1). With a continued decline in the popu- lation, a 1998 ASMFC amendment began a 40-year moratorium in order to protect 20 year classes of spawning females (ASMFC2). Currently, Atlan- tic sturgeon are a candidate species to be listed under the United States Endangered Species Act. Atlantic sturgeon use river, estua- rine, coastal, and oceanic environ- ments at different life stages but spend the majority of their lives in saltwater (Smith and Clugston, 1997). However, information on oceanic habi- tat use is lacking beyond evidence of broad-scale marine migrations and an exchange of populations among river systems based on tag recaptures (Dovel and Berggren, 1983) and com- mercial fisheries bycatch data (Stein et al., 2004a, 2004b). Fisheries-de- pendent data indicate that most At- 1 Atlantic States Marine Fisheries Com- mission (ASMFC). 1990. Fishery man- agement plan for Atlantic sturgeon. Fishery management report number 17, 85 p. ASMFC, Washington, D.C. 2 Atlantic States Marine Fisheries Com- mission (ASMFC). 1998. Amendment 1 to the interstate fishery management plan for Atlantic sturgeon, Fishery Manage- ment report 31, 59 p. ASMFC, Washing- ton, D.C. Dunton et at: Abundance and distribution of Aapenser oxyrinchus within the Northwest Atlantic Ocean 451 75°0'W 70°0'W 65°0'W Figure 1 Coverage area of the Maine-New Hampshire inshore bottom trawl survey (ME-NH), Massachusetts Division of Marine Fisheries bottom trawl survey (MADMF), New York bottom trawl survey (NYBTS), New Jersey Department of Environmental Protec- tion finfish survey (NJDEP), and the National Marine Fisheries Service bottom trawl surveys (NMFS). The area covered by the NMFS survey is represented by horizontal stripes. All other surveys are represented by shades of gray. lantic sturgeon inhabit shallow in- shore areas of the continental shelf (Stein et al., 2004a, 2004b). More re- cently, some long-term fishery-inde- pendent data have revealed that juve- nile Atlantic sturgeon use the inshore waters of North Carolina during the winter months (Laney et al., 2007). Additionally, there are a handful of reported cases of Atlantic sturgeon captured in deeper offshore areas (Ti- moshkin, 1968; Collins and Smith, 1997; Stein et al., 2004a, 2004b). Still, more information is needed to guide management towards the best mechanisms to protect the remaining Atlantic sturgeon. One contributing factor to the con- tinued decline of Atlantic sturgeon populations is incidental capture of juveniles in non-target marine fish- eries (Collins et al., 1996; Stein et al., 2004a). Most of the current by- catch mortality occurs in gill and drift net fisheries (Stein et al., 2004a; ASSRT3). Discard mortality from trawl fisheries is hard to esti- mate because few direct mortalities are observed. Mortality however may be very high due to delayed effects on cap- tured individuals (Davis, 2002; Broadhurst et al., 2006). Because Atlantic sturgeon do not reach maturity until 12-14 years of age and reproductive output increases later in life (Van Eenennaam and Doroshov, 1998), reducing mortality on juveniles is key to restoring depleted populations (Boreman, 1997). In order to adequately protect both juvenile and adult Atlantic sturgeon, marine distributional patterns must be identified such that essential habitat may be pro- tected. In this article we use data from five different oceanic fishery-independent surveys to reveal the sea- sonal distribution, abundance, and habitat use of Atlan- tic sturgeon along the Northwest Atlantic continental shelf from Cape Hatteras, NC, to the Gulf of Maine (GOM) (Fig. 1). Materials and methods We analyzed data from five fishery-independent sur- veys conducted by the following agencies: 1) National 3 Atlantic Sturgeon Status Review Team (ASSRT). 2007. Status review of Atlantic sturgeon ( Acipenser oxyrinchus oxyrin- chus), 174 p. Report to National Marine Fisheries Service, Northeast Regional Office, Gloucester, MA, 23 Feb 2007. National Oceanic and Atmospheric Science Administration, Washington D.C. Marine Fisheries Service (NMFS); 2) New Jersey Department of Environmental Protection (NJDEP); 3) Maine Department of Marine Resources and the New Hampshire Fish and Game Department (ME- NH); 4), Massachusetts Division of Marine Fisheries (MADMF); and 5) New York Bottom Trawl Survey (NYBTS) (Fig. 1). Catch per unit of effort (CPUE) was calculated (number of fish per tow) for each survey and depth (m). Depth (m), temperature (°C), and salinity (ppt) data were obtained from the NMFS, NJDEP, and NYBTS databases to estimate environmental prefer- ences. For all surveys, except the MADMF, depth was calculated as the average between the maximum and minimum values. Depth values used in the MADMF analysis are the depth at which the tow started. For all surveys, tows were analyzed for each season, which are defined as winter (21 Dec-20 Mar), spring (21 Mar-20 Jun), summer (21 Jun-20 Sept), and fall (21 Sep-20 Dec). Specifics of each survey are discussed in detail below. Because male and female Atlantic sturgeon mature at different size ranges (van Eenennaam and Doroshov, 452 Fishery Bulletin 108(4) 1998) and because we could not distinguish between gender, we applied female size at maturation to all individuals. Female maturation is reached at a total length of 197 cm (van Eenennaam and Doroshov, 1998). NMFS bottom trawl survey These surveys were conducted primarily by the research vessels Albatross IV and Delaware II where a Yankee 36 bottom trawl with a 1.27-cm mesh liner was towed for 30 minutes at 3.79 knots. Sampling was conducted during the day and night (Sosebee and Cadrin4). A total of 300-400 trawls were executed each season from the Gulf of Maine to just south of Cape Hatteras, NC (Fig. 1). Sampling for the NMFS fall survey began in 1963 and the waters of southern New England and the Gulf of Maine were sampled before tows were expanded to include inshore stations in 1973. The NMFS survey was further expanded to include spring samples in 1973. We also used some additional NMFS surveys that were con- ducted during the winters of 1964-66, 1972, 1978, 1981, and 1992-2007, and summers of 1977-81 and 1993-95. NJDEP finfish survey The NJDEP finfish survey began in 1988 and is con- ducted five times per year in April, June, August, Octo- ber, and January. A total of 186 tows are conducted each year (39 stations per trip for spring-fall months and 30 stations per trip for winter months). Sampling occurred from NY Harbor to the entrance of Delaware Bay, DE, from 8 to 30 m depth (Fig. 1). A depth-stratified random sampling design was used and a minimum of 10 tows were completed per depth interval (0-10 m, 10-20 m, and 20-30 m). The survey was conducted with a three- to-one two-seam trawl (25-m headrope, 30.5-m footrope) with 12-cm stretched mesh forward netting that tapered down to 8-cm stretched mesh rear netting that was lined with a 6.4-mm mesh codend liner. Tows were conducted at a speed of 3-3.5 knots for a duration of 20 minutes during daylight hours. ME-NH inshore bottom trawl survey This survey began in fall of 2000 and primarily covered the inshore waters of Maine and New Hampshire and a depth range of 9-150 m and distance up to 19.3 km offshore (in accordance with the 12-mile territorial limit) (Fig. 1). A total of 115 trawls were attempted each fall and spring. 100 stations were selected on the basis of a depth-stratified, random sampling design and of the 100 stations, 15 were fixed location stations. For this 4 Sosebee, K. A., and S. X. Cadrin. 2006. A historical perspective on the abundance and biomass of northeast demersal complex stocks from NMFS and Mas- sachusetts inshore bottom trawl surveys, 1963—2002. NEFSC Ref. Doc. 06-05, 200 p. Northeast Fisheries Science Center, National Marine Fisheries Service, Woods Hole Laboratory, 166 Water St., Woods Hole, MA 02543. survey a 57-70 modified shrimp trawl (17.37-m head rope, 21.34-m footrope) was used with 5.08-cm stretched mesh and a 2.54-cm stretched mesh liner in the codend. Tows were conducted for 20 minutes at 2. 2-2. 3 knots during daylight hours. MADMF bottom trawl survey Conducted during the spring and fall from 1978-2007, this bottom trawl survey encompassed the Massachu- setts inshore waters up to 5.6 km from the boundaries of New Hampshire and Rhode Island (Fig. 1). A % size North Atlantic two-seam otter trawl (head rope 11.9 m, footrope 15.5 m) with a 6.4-mm lined codend was towed at 2.5 knots for 20 min during daylight hours. The survey sampled 100 stations per year selected using a depth-stratified, random sampling design. New York bottom trawl surveys (NYBTS) The NY surveys consisted of two surveys — the New York young-of-the-year bluefish survey and the NY trawl survey for subadult Atlantic sturgeon. The sampling area encompassed the waters inshore of a depth of 30 m; the practical inshore limit was 8-10 m from Montauk Point to the entrance of NY Harbor (Fig. 1). For this survey a depth-stratified sampling design was used with strata based on the depth intervals 0-10 m, 10-20 m, and 20-30 m. Tows were randomly selected by using a random number generator and were conducted for a duration of 20 minutes at a tow speed of 3-3.5 knots during daylight hours. The net was a three-to-one two- seam trawl (25-m headrope, 30.6-m footrope) with for- ward netting of 12-cm stretched mesh tapering down to the rear netting of 8-cm stretched mesh and lined with a 6.0-mm mesh liner within the codend. Because exactly the same gear was used for the surveys, they were combined for the purpose of this analysis. Further differences between the two surveys are described below. The NY young-of-the-year bluefish survey was ini- tially restricted to the 10- and 20-m depth strata where 10 tows per depth stratum were completed for a total of 20 tows per cruise. Sampling took place June-October in 2005 and August-September in 2006. The survey was confined to the 10-m depth strata in September, October, and November of 2007 when 25, 24, and 27 tows were completed, respectively. For the NY trawl survey for subadult Atlantic stur- geon, a total of 10 cruises were conducted from Octo- ber 2005 through June 2007 with 30 tows per cruise distributed within the 10-, 20-, and 30-m depth strata. Sampling months were October, November, January, April, May, and June. A total of 10 tows were completed for each depth. In June 2007, 36 tows were confined to the 10-m depth stratum. Spatial analysis Atlantic sturgeon captures were mapped by season with ESRI® ArcGIS™, vers. 9.2 software (ESRI; Redlands, Dunton et al.: Abundance and distribution of Acipenser oxyrinchus within the Northwest Atlantic Ocean 453 CA). Map base layers were obtained from the United States Geological Survey Coastal and Marine Geology Program GIS catalogue. Atlantic sturgeon captures were plotted by using graduated symbols in the following categories: 1, 2, 3—4, 5—10, 11-14, and >15 Atlantic sturgeon per tow. Habitat preferences We estimated the habitat preference of Atlan- tic sturgeon by using the catch-weighted methods of Perry and Smith (1994) for cor- recting bias that arises in stratified sur- veys where sampling effort differs between strata. With this method, a comparison of a catch-weighted cumulative distribution of available (all habitat sampled) and occupied (habitat where Atlantic sturgeon were cap- tured) habitat was made and a randomization routine was used to estimate whether the occupied habitat was significantly different from available habitat. Habitat variables analyzed included temperature, dissolved oxygen, and salinity. The cumulative distribution function (cdf) of the environmental variable was calculated with the following function: 0.35 0.30 - I fe °'25 - Q. = 0.20 - 'o <5 | 0.15 - c LU 0.10 - Z3 CL O 0.05 - 0.00 T 1 1 T 1 1 1 1 1 NYBTS NJDEP ME-NH NMFS MADMF Figure 2 Catch per unit of effort (CPUE) for Atlantic sturgeon ( Acipenser oxyrinchus) during the New York bottom trawl survey (NYBTS), New Jersey Department of Environmental Protection finfish survey (NJDEP), Maine-New Hampshire inshore bottom trawl survey (ME-NH), National Marine Fisheries Service bottom trawl surveys (NMFS), and Massachusetts Division of Marine Fisheries bottom trawl survey (MADMF). , w f(t)= X H—KXhO’ (1) h , nh where W, n. the proportion of the survey in stratum /z; the number of tows in stratum h ; the habitat variable in tow i and stratum h; and an indicator function where | 1, if Hi^t 0, otherwise (2) The following function relates the catch weighted cdf to the habitat variable: *<*>=XX nh yst I(xlu), (3) Significance is determined by randomizing for 1000 trials the pairings of xhl and ( Wh/nh ) (yhi- yst)!yst ) and by dividing the number of trials that are greater than the test statistic by the total number of trials. Results The NYBTS had the highest CPUE (0.291 fish/tow), followed by the NJDEP finfish survey (0.072 fish/tow), ME-NH inshore bottom trawl survey (0.024 fish/tow), NMFS bottom trawl survey (0.004 fish/tow), and the MADMF bottom trawl survey (<0.001fish/tow), in the latter of which only one Atlantic sturgeon has ever been captured (Table 1; Fig. 2). The details of the CPUE by depth (Fig. 3) and seasonal distribution and abundance (Fig. 4-7) for each survey are reported in detail below. Total length of Atlantic sturgeon captured within the surveys ranged from 56 to 269 cm (mean=108 cm) (Table 2; Fig. 8). where yhi = the number of fish captured in tow i in stratum h ; and yst = the stratified mean abundance. The strength of the association is measured by the dif- ference between the available and occupied cdf: c|g(U-/'(U| = max XX— h i nh f — \ y -y , hi st V st J (4) NMFS bottom trawl survey A total of 107 Atlantic sturgeon were captured in 27,420 bottom trawls (Table 1). The depth distribution of com- pleted tows ranged from 5 to 542 m deep, and 5214 peak tows occurred between 20 and 40 m (Fig. 3A). CPUE of Atlantic sturgeon was highest for the 10 -m depth stra- tum (0.0273/tow) and decreased with each depth interval (Fig. 3A). A total of 71.30% of the Atlantic sturgeon were captured in 20 m or less and no individuals were captured in water deeper than 30 m (Fig. 3A). Atlantic 454 Fishery Bulletin 108(4) Table t Summary of the surveys effort and Atlantic sturgeon ( Acipenser oxyrinchus ) captures for the New York bottom trawl survey (NYBTS), New Jersey Department of Environmental Protection (NJDEP) finfish survey, National Marine Fisheries Service (NMFS) bottom trawl survey, Maine Department of Marine Resources and New Hampshire Fish and Game (ME-NH) inshore trawl survey, and Massachusetts Division of Marine Fisheries (MADMF) bottom trawl survey. Seasons are defined as winter (21 Dec-20 Mar), spring (21 Mar-20 Jun), summer (21 Jun-20 Sep), and fall (21 Sep-20 Dec). Survey Time period Total number of trawls completed Total number of Atlantic sturgeon captured Catch per unit of effort NYBTS 2005-07 512 149 0.291 fall 132 46 0.348 winter 59 4 0.068 spring 219 73 0.333 summer 102 26 0.255 NJDEP 1988-2007 3617 261 0.072 fall 769 74 0.096 winter 599 74 0.124 spring 1439 113 0.079 summer 810 0 0.000 NMFS 1973-2007 27,420 107 0.004 fall 11,919 26 0.002 winter 2563 12 0.005 spring 11,395 68 0.006 summer 1543 1 0.001 ME-NH 2000-06 1601 38 0.024 fall 773 31 0.040 spring 828 7 0.008 MADMF 1978-2007 5563 1 >0.001 spring 2874 1 >0.001 fall 2689 0 >0.001 Table 2 Mean, standard deviation, and range of total length (cm) of Atlantic sturgeon ( Acipenser oxyrinchus ) captured in the New York bottom trawl survey (NYBTS), New Jersey Department of Environmental Protection (NJDEP) fin- fish survey. National Marine Fisheries Service (NMFS) bottom trawl survey, Maine Department of Marine Resources and New Hampshire Fish and Game (ME- NH) inshore trawl survey, and Massachusetts Division of Marine Fisheries (MADMF) bottom trawl survey. Length information includes all recorded lengths over the dura- tion of the entire period of the above surveys. Mean total length ±standard Survey deviation (cm) Range (cm) NYBTS 112.01 ±27.75 72-215 NJDEP 103.89 ±32.13 52-248 NMFS 113.87 ±40.18 51-269 ME-NH 115.4 ±19.39 76-152 MADMF 78 ±0 — sturgeon were captured during all seasons but were most abundant during the spring, with an average CPUE of 0.006 fish/tow, followed by winter (0.005 fish/tow), fall (0.002 fish/tow), and summer (0.001 fish/tow) (Table 1). In the spring, 70.59% of Atlantic sturgeon were cap- tured in Virginia (VA) and NC waters and 23.53% were captured in NY and NJ. One Atlantic sturgeon was captured south of Cape Hatteras and one was captured offshore of northern MA. During winter months cap- tures were evenly distributed from NJ to NC. A total of 42.30% (11 fish) of fall captures occurred off Long Island, NY, whereas 30.76% (8 fish) occurred in the mouth of Delaware Bay, Delaware (DE). In addition three fish were captured in NJ, one fish south of Cape Hatteras, and one fish near Cape Cod, MA. Only one Atlantic sturgeon was captured during this survey in the summer months in NY waters off of Long Island. NJDEP finfish survey A total of 261 Atlantic sturgeon were captured within 3617 bottom trawls from 1988 through 2007 (Table 1) at all depths sampled (Fig. 3B). Tow distribution ranged Dunton et al.: Abundance and distribution of Acipenser oxyrinchus within the Northwest Atlantic Ocean 455 0.030 0.025 0.020 0.015 0.010 0.005 0.000 A (NMFS) 0.160 B (NIDEP) 0160 < 0.140 0.140 0.120 0.120 0.100 0.100 0.080 0.080 0.060 0.060 0.040 0.040 0.020 , 0.020 .-11 , , , , , - — . 0.000 — . — . — . — . — , — - — < o.ooo 0 50 1 00 1 50 200 250 300 350 400 0 50 100 150 200 250 300 350 400 (ME-NH) .il.n. 0 50 1 00 1 50 200 250 300 350 400 Depth >. o c 15 38°0'N 76°0'W 72°0'W Figure 4 Number of captures of Atlantic sturgeon ( Acipenser oxyrinchus ) from all surveys during spring months. Circle size corresponds to total number of Atlantic sturgeon captured at a given location (insert A). Locations of all tows can be seen in insert B. the total Atlantic sturgeon captured, 64% (48 fish) were captured off northern NJ. Captures within the spring occurred along the entire coast of NJ and 44.2% of the captures occurred in Sandy Hook, NJ. ME-NH inshore bottom trawl survey A total of 38 Atlantic sturgeon were captured in a total of 1601 bottom trawls from 2001 through 2006 (Table 1). Sampling depths ranged from 10 to 200 m, and three defined peaks in sampling effort occurred at 30 m, 65 m, and 90 m (Fig. 3C). All Atlantic sturgeon were captured between 15 and 90 m depth (Fig. 30 and 36 of the 38 Atlantic sturgeon were captured near the Kennebec estuarine complex (Fig. 9A). Two additional Atlantic sturgeon were captured south of the Kennebec River, closer to the Saco River. MADMF bottom trawl survey Only one Atlantic sturgeon was captured in a total of 5563 bottom trawls (Table 1). Sampling depths ranged from 4 to 86 m, and a peak in sampling effort occurred at a depth of 20 m (Fig. 3D). The only Atlantic sturgeon captured was collected during the spring at a depth of 41 m. NYBTS A total of 149 Atlantic sturgeon were captured in 512 random stratified tows (Table 1). Sampling depths ranged from 5 to 35 m and a peak in sampling effort occurred at a depth of 15 m (Fig. 3E). Atlantic sturgeon were cap- tured within all months sampled; however, no Atlantic sturgeon were captured deeper than 20 m. A total of 85% of all Atlantic sturgeon were captured between 5 and 10 m with a mean CPUE of 1.34 (Fig. 3E). CPUE was high- est during the fall (0.35 fish/tow) followed by spring (0.33 fish/tow) and summer (0.26 fish/tow) and was lowest during the winter (0.07 fish/tow) (Table 1). Of the 149 Atlantic sturgeon captured, 51% were collected off the western coast of Long Island, 30% were captured off cen- tral Long Island, and only one was captured off the east end of Long Island. During the spring, Atlantic sturgeon were captured along the entire coast of Long Island, NY, but 57% were captured off western Long Island, specifically Rockaway, NY. The Rockaway region was also an important area during the fall, accounting for Dunton et al.: Abundance and distribution of Acipenser oxyrinchus within the Northwest Atlantic Ocean 457 76°0 W 72°0'W Figure 5 Number of captures of Atlantic sturgeon ( Acipenser oxyrinchus) from all surveys during winter months. Circle size corresponds to total number of Atlantic sturgeon captured at a given location (insert A). Locations of all tows can be seen in insert B. 70% of the catches occurring within this region. Twenty- six Atlantic sturgeon were captured in the summer months; 99% were captured in western-central Long Island, NY, and only one was captured along the east end of Long Island. During the winter, all Atlantic stur- geon were captured off the western end of Long Island. Habitat preferences Hydrographic variables and distributions of Atlantic sturgeon were compared only for the NMFS bottom trawl survey, NJDEP finfish survey, and for NYBTS for the spring and fall seasons because these contained sufficient Atlantic sturgeon capture data to perform the analyses. The depths (habitat) occupied by Atlantic sturgeon was significantly different from the available depths in the NMFS survey and NYBTS for both the spring and fall surveys and the NJDEP spring survey (Tables 3 and 4). Atlantic sturgeon occupied areas with significantly different temperatures compared to available habitat in the NYBTS spring and NMFS fall survey, as well as areas of significantly different salini- ties in the NMFS fall and spring surveys and NJDEP spring survey (Table 3). Survey-specific cumulative distribution functions for available and occupied habits Table 3 P-values from the analysis of habitat preference of Atlan- tic sturgeon (Acipenser oxyrinchus ) by season in the northwestern Atlantic Ocean. The fish were captured in the National Marine Fisheries Service (NMFS) bottom trawl survey, New Jersey Department of Environmental Protection (NJDEP) finfish survey, and New York bottom trawl survey (NYBTS). Bold font indicates a significant difference (P<0.01) between Atlantic sturgeon habitat preference and the available habitat. Season Survey Depth Temperature Salinity fall NMFS <0.005 <0.005 <0.005 fall NJDEP 0.129 0.173 0.273 fall NYBTS <0.005 0.518 0.530 spring NMFS <0.005 0.355 0.001 spring NJDEP <0.005 0.173 <0.005 spring NYBTS <0.005 0.001 0.084 with depth, salinity, and temperature profiles are shown in Figure 10 and median values and 95% confidence intervals are listed in Table 4. Where significant differ- 458 Fishery Bulletin 108(4) 76°0'W 72°0'W 42°0'N 38°0'N - -42°0'N - 38°0'N 76°0'W 72°0'W Figure 6 Number of captures of Atlantic sturgeon ( Acipenser oxyrinchus) from all surveys during fall months. Circle size corresponds to total number of Atlantic sturgeon captured at a given location (insert A). Locations of all tows can be seen in insert B. ences in depth occurred, Atlantic sturgeon were always found in shallower water than in potentially available habitat (Table 4, Fig. 10). Salinities of occupied areas were less than those of available habitat in all surveys, although only the NMFS fall and spring survey and NJDEP spring survey had significant differences (Table 4, Fig. 10). In two circumstances the temperature of occupied habitat was significantly warmer than that of available habitat, whereas temperature for the other seasons and surveys showed no trend (Table 4, Fig. 10). Discussion A majority of the Atlantic sturgeon captured along the continental shelf from ME to NC were juveniles aggre- gating in specific locations around the mouths of estua- rine complexes and along narrow dispersal corridors in shallow water (<20 m) from Cape Hatteras (NC) to the southern tip of Long Island (NY). The highest catches occurred within the NY Bight in water 10-15 m deep, particularly during the spring and fall. Few sturgeon were captured north of MA. Little work has been done to describe the marine habitat distribution and habi- tat preference of Atlantic sturgeon, but similar coast- wide, shallow (with respect to regional bathymetry) marine distributions have been shown for green stur- geon ( Acipenser medirostris) (Erickson and Hightower, 2007) and Gulf sturgeon ( Acipenser oxyrinchus desotoi) (Edwards et al., 2007; Ross et al., 2009). The shallow, coast-wide habitat identified within this study is also consistent with Atlantic sturgeon bycatch data (Stein et al., 2004a, 2004b). Our comprehensive analysis of a coast-wide collection of surveys revealed the area between the NY Bight to VA as a region of overwinter- ing habitat for juvenile Atlantic sturgeon. This finding agrees with that of Laney et al. (2007), who found the coastal waters off NC and VA to be important overwin- tering habitat for Atlantic sturgeon. Atlantic sturgeon that originated from the Hudson River represented 43.5% of those in the NC overwintering habitat (Laney et al., 2007) a percentage that agrees with Dovel and Berggren’s (1983) tagging data that revealed a south- erly movement of Atlantic sturgeon from the Hudson River. In addition to Laney et al. (2007), there have been further reports of Atlantic sturgeon in marine waters off the coast of South Carolina during winter months (Collins and Smith, 1997). The identification of the NY Bight as an important overwintering area has not been widely reported; therefore determining the Dunton et al.: Abundance and distribution of Acipenser oxynnchus within the Northwest Atlantic Ocean 459 76°0'W 72°0'W Figure 7 Number of captures of Atlantic sturgeon ( Acipenser oxyrinchus ) from all surveys during summer months. Circle size corresponds to total number of Atlantic sturgeon captured at a given location (insert A). Locations of all tows can be seen in insert B. genetic makeup of the sturgeon in this area would add important information on Atlantic sturgeon demographics and movements. Atlantic sturgeon had a coast-wide distribu- tion during the spring and fall, and southerly and centrally located distributions during the winter and summer, respectively. These results corroborate tagging data that indi- cate that Atlantic sturgeon undergo large- scale southerly fall migrations and north- erly spring migrations (Dovel and Berggren, 1983). Catches varied by season, but were greatest during the fall and spring months. Because of the strong seasonal movements of Atlantic sturgeon, the timing of surveys is critical for observing movement patterns. Interaction of Atlantic sturgeon abundance with temporal and spatial variability in sampling effort Some of the variation in distribution and abundance of Atlantic sturgeon can be explained by temporal and spatial differ- Total length (cm) Figure 8 Total length distribution for Atlantic sturgeon (Acipenser oxyrin- chus) captured in all surveys combined. 460 Fishery Bulletin 108(4) ences in sampling effort. Stein et al. (2004a) reported that MA ports have one of the highest cumulative catches of Atlan- tic sturgeon. This high rate contrasts with that for the MADMF bottom trawl survey, during which virtually no Atlantic sturgeon were captured. The discrepancy between reports of Atlantic sturgeon in MA waters likely comes from the timing of sampling. Stein et al. (2004a) showed the highest bycatch rates in June and Novem- ber for bottom trawl fisheries, whereas the MADMF survey took place during May and September. Any aggregations and dispersal of Atlantic sturgeon within MA marine waters may occur at spatial and temporal scales that are missed by the survey. The absence of Atlantic sturgeon during the MADMF survey does, however, indicate lower abundance within the area surveyed during comparable time frames because Atlantic sturgeon are captured at relatively high rates by other surveys during this period. More work should be done to monitor Atlantic sturgeon habitat during other months not typically sampled by the MADMF survey because it is pos- sible that Atlantic sturgeon are present in higher concentrations during months that are not routinely sampled. The NMFS survey missed critical ar- eas for Atlantic sturgeon because inshore areas in certain regions could not be sampled. Such areas include important overwintering habitat identified within this study in NY waters and by Laney et al. (2007) in VA and NC waters, in addi- tion to critical habitat within the GOM. The ME-NH inshore bottom trawl survey was used to identify the Kennebec es- tuarine complex as an important concentration area for Atlantic sturgeon within the GOM re- gion because shallower areas are sampled with this survey. During additional inshore surveys, such as the Northeast Fisheries Sciences Center (NEFSC) industry-based surveys for cod ( Gadus morhua) and yellowtail ( Limanda ferruginea), Atlantic sturgeon have been captured between the Saco and Kennebec rivers in fall, winter, and spring (Fig. 9A; W. Kramer, personal com- mun.5). Stein et al. (2004a, 2004b) also showed that Atlantic sturgeon are captured as bycatch within this region in sink gillnets. The depth distribution of Atlantic sturgeon within the GOM was deeper than that for the other coast-wide captures, but similar to those reported for green 70°20'W 70°0'W 69°40'W 74°0'W 73°50'W 73°40'W 5 Kramer, William. 2009. NOAA Fisheries Service, Ecosystems Survey Branch, 166 Water St., Woodshole, MA 02543. Figure 9 Number of captures of Atlantic sturgeon ( Acipenser oxyrin- clius) determined from (A) the Gulf of Maine from the Maine- New Hampshire bottom trawl surveys (gray circles) and from the Northeast Fisheries Science Center industry-based surveys for cod (Gadus morhua) and yellowtail (Limanda ferruginea) (black circles) and (B) Sandy Hook, NJ and Rockaway NY, including all captures from the National Marine Fisheries Service bottom trawl survey. New Jersey Department of Environmental Protection finfish survey, and New York bottom trawl survey. Dotted lines in both panels represent suggested closed areas for habitat protec- tion. Note different scales in figure. sturgeon (Erickson and Hightower, 2007) in that both species occupied areas of shallow depth in relation to the bathymetric characteristics of the region. There has not been sufficient inshore trawling conducted during the winter and summer to validate whether the GOM is important year-round habitat. Dunton et al.: Abundance and distribution of Acipenser oxyrinchus within the Northwest Atlantic Ocean 461 Table 4 Median and 95% confidence intervals for available and occupied habitat of Atlantic sturgeon (Acipenser' oxyrinchus ) for depth, temperature, and salinity for the fall and spring National Marine Fisheries Service bottom trawl survey (NMFS), New Jersey Department of Environmental Protection finfish survey (NJDEP), and New York bottom trawl survey (NYBTS) where “available habitat” represents all habitat sampled and “occupied habitat” represents habitat where Atlantic sturgeon were captured. Season Parameter Survey Median 95% confidence interval Fall Depth (m) NMFS available habitat 76.0 15.5-260.5 NMFS occupied habitat 18.0 10.0-25.0 NJDEP available habitat 19.0 8.0-27.0 NJDEP occupied habitat 16.0 7.0-22.0 NYBTS available habitat 23.8 9.5-30.3 NYBTS occupied habitat 10.7 9.0-17.0 Temperature (°C) NMFS available habitat 10.6 5.9-22.5 NMFS occupied habitat 18.9 13.3-23.3 NJDEP available habitat 15.8 12.3-18.6 NJDEP occupied habitat 14.8 13.3-17.6 NYBTS available habitat 15.3 13.5-19.3 NYBTS occupied habitat 16.8 13.6-19.9 Salinity (ppt) NMFS available habitat 33.1 31.0-35.4 NMFS occupied habitat 31.6 29.3-32.0 NJDEP available habitat 32.0 29.6-33.5 NJDEP occupied habitat 31.5 29.5-33.1 NYBTS available habitat 31.4 30.1-32.9 NYBTS occupied habitat 31.3 29.4-31.8 Spring Depth (m) NMFS available habitat 76.0 16.0-259.0 NMFS occupied habitat 18.0 8.0-27.0 NJDEP available habitat 19.0 7.5-27.0 NJDEP occupied habitat 12.0 7.0-18.0 NYBTS available habitat 22.4 9.9-29.7 NYBTS occupied habitat 9.9 9.9-13.9 Temperature (°C) NMFS available habitat 6.0 3.4-12.5 NMFS occupied habitat 6.4 3.2-15.0 NJDEP available habitat 9.1 4.9-18.8 NJDEP occupied habitat 11.0 5.7-19.0 NYBTS available habitat 9.4 5.1-14.6 NYBTS occupied habitat 11.1 5.7-13.9 Salinity (ppt) NMFS available habitat 33.2 31.4-35.4 NMFS occupied habitat 32.0 27.0-32.8 NJDEP available habitat 32.0 30.0-34.0 NJDEP occupied habitat 30.0 28.8-35.0 NYBTS available habitat 31.6 30.2-33.1 NYBTS occupied habitat 30.9 29.9-32.3 Although the NMFS survey covered the entire conti- nental shelf, no fish were captured deeper than 30 m. However, Atlantic sturgeon of unknown size have been captured in deeper waters (>100 m) on the continental shelf as bycatch in gillnet fisheries (Stein et al. 2004b; ASMFC6). Additionally, there have only been two re- corded trawl captures of an Atlantic sturgeon in deep waters; one mature Atlantic sturgeon (225 cm) was cap- tured in the Hudson canyon in water 110 m deep off NY and NJ and another was captured in Wilmington Can- yon, 113 km southeast of Atlantic City, NJ (Timoshkin, 1968). The lack of trawl-caught fish on the continental shelf may be a result of either a gap in timing of the sampling and Atlantic sturgeon migrations on or off the shelf, a function of gear selectivity towards smaller 6 ASMFC (Atlantic States Marine Fisheries Commission). 2007. Estimation of Atlantic sturgeon bycatch in coastal Atlantic commercial fisheries of New England and the Mid-Atlantic, 95 p. ASMFC, Washington D.C. 462 Fishery Bulletin 108(4) A (NMFS) B (NJDEP) C (NYBTS) Depth (m) Figure 10 Cumulative distribution functions for available habitat and habitat occupied by Atlantic stur- geon ( Acipenser oxyrinchus) in the fall and spring surveys for (A) National Marine Fisheries Service bottom trawl surveys (NMFS), (B) New Jersey Department of Environmental Protection finfish survey (NJDEP), and (C) New York bottom trawl survey (NYBTS) for depth (m), salinity (ppt), and temperature (°C). Solid lines indicate habitat occupied by A. oxyrinchus (gray=fall; black=spring) and dashed lines indicate available habitat (dashed gray=fall; dashed black= spring). fish, or simply a scarcity of Atlantic sturgeon. Either a substantial increase in trawl survey effort or the use of different gears, such as gillnets, may be required in order to capture Atlantic sturgeon along the shelf. Essential fish habitat The Magnuson-Stevens Fishery Conservation and Man- agement Act requires identification of Essential Fish Habitat (EFH), defined as waters or substrate used for spawning, breeding, feeding or growth to maturity, in order to minimize adverse effects of fishing and to promote conservation and enhancement of such habitat for particular species. Unfortunately, EFH can only be defined for federally managed species and does not include species such as Atlantic sturgeon which are managed by regional fishery management councils. Atlantic sturgeon is a current candidate species for listing under the US Endangered Species Act, and if listed, the identification of critical habitat necessary to recover the species will be required. The identification of critical habitat for listed species is mandatory and is defined as all areas essential to the conservation of the species. Without EFH or critical habitat desig- nation, habitat degradation and incidental mortality within critical areas will continue to hinder population recovery. Our analysis of habitat preferences indicated that depth was the primary environmental characteristic defining the Atlantic sturgeon distribution. Thus, es- Dunton et al.: Abundance and distribution of Acipenser oxyrmchus within the Northwest Atlantic Ocean 463 sential habitat for juvenile marine migrant Atlantic sturgeon can broadly be defined as coastal waters <20 m depth, and it is concentrated in areas adjacent to es- tuaries such as the Hudson River-NY Bight, Delaware Bay, Chesapeake Bay, Cape Hatteras, and Kennebec River. This narrow band of shallow water appears to represent an important habitat corridor and potential migration path. There are likely additional hotspots along the migration corridor, but greater temporal and spatial sampling effort is required to identify them. Other authors have reported concentrations of Atlantic sturgeon in Long Island Sound (Bain et al., 2000; Savoy and Pacileo, 2003) and NC (Laney et al., 2007), and Stein et al. (2004a) reported several concentrations of Atlantic sturgeon in Massachusetts Bay, RI, NJ, and DE. However, Stein et al. (2004a) used bycatch data in areas where captures were lowest during the summer months while the fishing rates were highest. However, this change in fishing effort may influence the observed distributions. The reason(s) for aggregations of Atlantic sturgeon migrants are not understood, nor are their movements to and from aggregation areas. Concentrations identified by Stein et al. (2004b) led the authors to suggest that temperature, bathymetry, geomorphic formations, food habits, and the sampling gear type used may contrib- ute to observed movements and aggregation of Atlantic sturgeon. Complex water circulation patterns are also a potential reason for observed concentrations of At- lantic sturgeon (Wilk and Silverman, 1976; Savoy and Pacileo, 2003). Hatin et al. (2002) found that Atlantic sturgeon concentrated within the St. Lawrence estu- ary had large numbers of nematodes and oligochaetes within their stomachs, which would indicate that these habitats are feeding areas. Known seasonal migra- tions often involve energetic demands related to food availability, environmental factors, and reproductive activity (Roff, 2002). Because the majority of captures are juveniles, reproductive activity is not a likely cause for movement, although causal mechanisms influencing traits under selection are difficult to identify because life-history stages are often linked through long-term fitness (Taborsky, 2006). We hypothesize that migra- tions are depth restricted and aggregations are related to food availability, and that seasonal cues, temperature in particular, drive movement. Current and future management of Atlantic sturgeon Current knowledge indicates that the majority of Atlan- tic sturgeon populations have been extirpated and that the Hudson River stock is one of the largest remaining populations (Waldman et al., 1996; van Eenennaam et al., 1998; Savoy and Pacileo, 2002; Secor et al., 2002). Three fishery management tools commonly used to help restore depleted populations are 1) minimum size limits, 2) temporary closures of the fishery, and 3) marine reserves (Nowlis, 2000). Management of Atlantic stur- geon has been accomplished by minimum size limits since the early 1990s, followed directly by a 40-year complete closure of the fishery beginning in 1998. Cur- rently, after 10 years of the fishery closure, recruitment within the Hudson River still remains at historic lows (Kahnle et al., 2007). Because previous Atlantic sturgeon management has not resulted in significant improvements to popula- tions, recovery efforts should now focus on establish- ing marine reserves or implementing area closures to protect essential habitat and to reduce fishing mortality on juveniles (Collins et al., 2000). Specifi- cally, Sandy Hook (NJ), Rockaway (NY), and Ken- nebec (ME), which are hotspots of Atlantic sturgeon captures, as identified by this study, should be pro- tected. Although sturgeon are not as abundant in the Kennebec region in ME as in NY and NJ waters, this region represents a unique hotspot. It is of particular importance because Atlantic sturgeon captured in ME river systems have been shown to represent a separate discrete population segment (Grunwald et al., 2008). The genetic origins of the Atlantic sturgeon captured within marine waters of ME are unknown, but they are likely to originate from multiple stocks. Because of the proximity of ME river systems, it is probable that the majority of these Atlantic sturgeon are part of this discrete population segment. If our recommended habitat protection were to occur, the total amount of closed area within these locations would be relatively small — totaling 85.47 km2 within NJ (Fig. 9A), 106.19 km2 within NY (Fig. 9A), and 209.79 km2 within ME (Fig. 9B). In addition, although Atlantic sturgeon are highly migratory, primary juvenile habitat and migra- tions are limited to narrow corridors in waters less than 20 m deep. The presence of Atlantic sturgeon in such narrow bands of water indicates a seasonal or permanent closure to gillnet and trawl fisheries could be successful. By focusing immediate efforts on the protection of these hotspots and corridor pathways, bycatch mortality will be reduced effectively through protection of habitat. Further efforts should also be made to protect important areas within other systems and to conserve the several discrete population seg- ments defined by ASSRT3 and Grunwald et al. (2008) and to promote genetic diversity among Atlantic stur- geon populations. By understanding the time periods of localized ag- gregations and movements of Atlantic sturgeon, plans could be developed that minimize the extent and length of closures that are concentrated within narrow cor- ridors. Some states already restrict inshore trawling which limits fishery interactions with Atlantic stur- geon, such as NJ (3.22 km limit), MD (1.61 km limit), DE (no trawling), and NY (various no trawl zones in marine waters). Any spatial closures require proper enforcement and substantial community-level support for successful implementation (Sumaila et al.. 2000). Although broad-scale movement patterns are becoming clearer, work is required to understand the finer scale movements of Atlantic sturgeon such that any spa- tial management plans could be minimized while still achieving adequate protection. Current plans toward 464 Fishery Bulletin 108(4) understanding finer-scale movements are aided by co- operative efforts such as those of the Atlantic Coopera- tive Telemetry (ACT) network, which is a large scale collaborative telemetry network of -30 groups from Maine to South Carolina (D. Fox and T. Savoy, personal commun.7'8). Such coordinated efforts are steps in the right direction for species conservation. Once fine-scale movements are understood, in particular for aggrega- tion areas, fishery managers will be better informed as to how to limit interactions between fisheries and the near-endangered Atlantic sturgeon while minimiz- ing economic impacts. Improving estimates of fishery bycatch mortality would be of enormous value, in par- ticular if these estimates included a spatial perspective. Regardless of the outcome of the current consideration of Atlantic sturgeon for listing under the endangered species act, a coordinated effort among academic, fed- eral, state, and local institutions will be required to conserve this ancient species. Acknowledgments We would like to thank W. Kramer (National Oceanic and Atmospheric Administration, National Marine Fish- eries Service), D. Byrne (New Jersey Department of Environmental Protection), J. Sowles (Maine Depart- ment of Marine Resources), and J. King (Massachusetts Division of Marine Fisheries) for providing the data sets used in this study. We would like to thank M. Wiggins for survey design and technical support and Captain S. Cluett and the crew of the RV Seawolf for helping collect data in New York waters. Funding for the NY trawl surveys was provided by State Wildlife Grant award no. 910245 and NOAA research grant no. NA07NMF4550320. The Steven Berkeley Fellowship award from the American Fisheries Society further supported the research of K. Dunton. Literature cited Bain, M. B., N. Haley, J. R. Waldman, and K. Arend. 2000. Harvest and habitats of Atlantic sturgeon Acipenser oxyrinclius Mitchill, 1815 in the Hudson River estu- ary: lessons for sturgeon conservation. Bol. Inst. Esp. Oceanogr. 16:43-53. Boreman, J. 1997. Sensitivity of North American sturgeons and paddlefish to fishing mortality. Environ. Biol. Fishes 48:399-405. Broadhurst M. K., P. Suuronen, and A. Hulme. 2006. Estimating collateral mortality from towed fishing gear. Fish Fish. 7:180-218. 7 Fox, Dwayne. 2009. Department of Agriculture and Natural Resources, Delaware State Univ., 1200 North Dupont Hwy., Dover, DE 19901. 8 Savoy, Tom. 2009. Connecticut Department of Environ- mental Protection, Marine Fisheries Division, P.O. Box 719, Old Lyme, CT 06371-0719. Collins, M. R., S. G. Rogers, and T. I. J. Smith. 1996. Bycatch of sturgeons along the Southern Atlantic coast of the USA. N. Am. J. Fish. Manag. 16:24-29. Collins, M. R., and T. I. J. Smith. 1997. Distributions of shortnose and Atlantic sturgeons in South Carolina. N. Am. J. Fish. Manag. 17:995-1000. Collins, M. R., S. G. Rodgers, T. I. J. Smith, and M. L. Moser 2000. Primary factors affecting sturgeon populations in the southeastern United States: fishing mortality and degradation of essential habitat. Bull. Mar. Sci. 66:917-928. Davis, M. W. 2002. Key principles for understanding fish bycatch dis- card mortality. Can. J. Fish. Aquat. Sci. 59:1834-1843. Dovel, W. L., and T. J. Berggren. 1983. Atlantic sturgeon of the Hudson Estuary, New York. NY Fish Game J. 30:140-172. Edwards, R. E., F. M. Parauka, and K. J. Sulak. 2007. New insights into marine migration and winter habitat of Gulf sturgeon. In Anadromous sturgeons: habitats, threats, and management (J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron , eds.), p. 183-196. Am. Fish. Soc. Symp. 56, Bethesda, MD. Erickson, D. L., and J. E. Hightower. 2007. Oceanic distribution and behavior of green stur- geon. In Anadromous sturgeons: habitats, threats, and management (J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron, eds.), p. 197-211. Am. Fish. Soc. Symp. 56, Bethesda, MD. Grunwald, C., L. Maceda, J. Waldman, J. Stabile, and I. Wirgin. 2008. Conservation of Atlantic sturgeon Acipenser oxy- rinchus: delineation of stock structure and distinct population segments. Conserv. Genet. 9:1111-1124. Hatin, D., R. Fortin, and F. Caron. 2002. Movements and aggregation areas of adult Atlantic sturgeon ( Acipenser oxyrinchus) in the St. Lawrence River estuary, Quebec, Canada. J. Appl. Icthyol. 18:586-594. Kahnle, A. W., K. A. Hattala, K. A. and McKown. 2007. Status of Atlantic sturgeon in the Hudson River Estuary, New York, USA. In Anadromous sturgeons: habitats, threats, and management (J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron , eds.) p. 347-363. Am. Fish. Soc. Symp. 56, Bethesda, MD. Laney, R. W., J. E. Hightower, B. R. Versak, M. F. Mangold, W. W. Cole Jr., and S. E. Winslow. 2007. Distribution, habitat use, and size of Atlantic sturgeon captured during cooperative winter tagging cruises, 1988-2006. In Anadromous sturgeons: habi- tats, threats, and management (J. Munro, D. Hatin, J. E. Hightower, K. McKown, K. J. Sulak, A. W. Kahnle, and F. Caron, eds.) p. 167-182. Am. Fish. Soc. Symp. 56, Bethesda, MD. Nowlis, J. S. 2000. Short- and long- term effects of three fishery- management tools on depleted populations. Bull. Mar. Sci. 66:651-662. Perry, R. I., and S. J. Smith. 1994. Identifying habitat associations of marine fishes using survey data: an application to the Northwest Atlantic. Can. J. Fish. Aquat. Sci. 51:589-602. Roff, D. 2002. Life history evolution, 527 p. Sinauer Associates, Inc., Sunderland, MA. Dunton et al.: Abundance and distribution of Acipenser oxyrinchus within the Northwest Atlantic Ocean 465 Ross, S. T., W. T. Slack, R. J. Heise, M. A. Dugo, H. Rogillio, B. R. Bowen, P. Mickle, and R. W. Heard. 2009. Estuarine and coastal habitat use of Gulf sturgeon ( Acipenser oxyrinchus destoi ) in the North Central Gulf of Mexico. Estuar. Coasts 32:360-374. Savoy, T., and D. Pacileo. 2003. Movements and important habitats of subadult Atlantic sturgeon in Connecticut waters. Trans. Am. Fish. Soc. 132:1-8. Secor, D. H., P. J. Anders, W. Van Winkle, and D. A. Dixon. 2002. Can we study sturgeons to extinction? What we do and don’t know about the conservation of North American sturgeons. In Biology, management, and protection of North American sturgeon (W. Van Winkle, P. J. Anders, D. H. Secor, and D. A. Dixon, editors, eds.), p. 3-10. Am. Fish. Soc. Symp. 28, Bethesda, MD. Smith, T. I. J., and J. P. Clugston. 1997. Status and management of Atlantic sturgeon, Acipenser oxyrinchus, in North America. Environ. Biol. Fishes 48:335-346. Stein, A. B., K. B. Friedland, and M. Sutherland. 2004a. Atlantic sturgeon marine bycatch mortality on the continental shelf of the northeastern United States. N. Am. J. Fish. Manag. 24:171-183. 2004b. Atlantic sturgeon marine distribution and hab- itat use along the northeastern coast of the United States. Trans. Am. Fish. Soc. 133:527-537. Sumaila, U. R., S. Guenette, J. Alder, and R. Chuenpagdee. 2000. Addressing ecosystem effects of fishing using marine protected areas. ICES J. Mar. Sci. 57:752-760. Taborsky, B. 2006. The influence of juvenile and adult environments on life-history trajectories. Proc. R. Soc. London, Ser. B 273:741-750. Timoshkin, V. P. 1968. Atlantic sturgeon (Acipenser sturio L.) caught at sea. Prob. Ichthyol. 8(4): 598 . van Eenennaam, J. P, and S. I. Doroshov. 1998. Effects of age and body size on gonadal develop- ment of Atlantic sturgeon. J. Fish Biol. 53:624—637. Waldman, J. R., J. T. Hart, and I. Wirgin. 1996. Stock composition of the New York Bight Atlantic sturgeon fishery based on analysis of mitochondrial DNA. Trans. Am. Fish. Soc. 125:364-371. Wilk, S. J., and M. J. Silverman. 1976. Summer benthic fish fauna of Sandy Hook Bay, New Jersey. NOAA Tech. Rep. NMFS SSRF-698, 16 p. 466 Day and night abundance, distribution, and activity patterns of demersal fishes on Heceta Bank, Oregon Ted D. Hart' Julia E. R. Clemons (contact author)2 W. Waldo Wakefield2 Selina S. Heppell' Email address for contact author: Julia.Clemons@noaa.gov 1 Oregon State University Department of Fisheries and Wildlife 104 Nash Hall Corvallis, Oregon 97331 Present address for Ted D. Hart: Environmental Science and Management Portland State University 1719 SW 10th Avenue, SB2 Room 218 Portland, Oregon 97201 2 Northwest Fisheries Science Center Fishery Resource Analysis and Monitoring Division National Marine Fisheries Service National Oceanic and Atmospheric Administration 2032 SE OSU Drive Newport, Oregon 97365 Abstract — Most shallow-dwelling tropical marine fishes exhibit differ- ent activity patterns during the day and night but show similar transition behavior among habitat sites despite the dissimilar assemblages of the species. However, changes in species abundance, distribution, and activ- ity patterns have only rarely been examined in temperate deepwater habitats during the day and night, where day-to-night differences in light intensity are extremely slight. Direct- observation surveys were conducted over several depths and habitat types on Heceta Bank, the largest rocky bank off the Oregon coast. Day and night fish community composition, relative density, and activity levels were compared by using videotape footage from a remotely operated vehicle (ROV) operated along paired transects. Habitat-specific abundance and activity were determined for 31 taxa or groups. General patterns observed were similar to shallow temperate day and night studies, with an overall increase in the abun- dance and activity of fishes during the day than at night, particularly in shallower cobble, boulder, and rock ridge habitats. Smaller school- ing rockfishes ( Sebastes spp.) were more abundant and active in day than in night transects, and sharp- chin (S. zacentrus ) and harlequin (S. variegatus) rockfish were significantly more abundant in night transects. Most taxa, however, did not exhibit distinct diurnal or nocturnal activ- ity patterns. Rosethorn rockfish (S. helvomaculatus) and hagfishes (Epta- tretus spp.) showed the clearest diur- nal and nocturnal activity patterns, respectively. Because day and night distributions and activity patterns in demersal fishes are likely to influence both catchability and observability in bottom trawl and direct-count in situ surveys, the patterns observed in the current study should be considered for survey design and interpretation. Manuscript submitted 30 September 2009. Manuscript accepted 24 August 2010. Fish. Bull. 108:466-477 (2010). The views and opinions expressed or implied in this article are those of the author (or authors) and do not necessarily reflect the position of the National Marine Fisheries Service, NOAA. For surveys of groundfish, swept-area trawls are generally conducted and direct-counts from video cameras are recorded during daylight hours (Gunderson, 1993; Adams et al., 1995; Yoklavich et al., 2000; Wakefield et al., 2005). If activity patterns vary among species during the day versus the night, conclusions about rela- tive abundance, community composi- tion, and habitat affiliations could be incomplete or biased. It is not known whether diel activity and abundance patterns in fishes commonly found in shallow temperate and tropical areas are similar for fishes inhabit- ing deeper temperate areas along the west coast of North America, where diel changes in light levels are subtle. We used a repeat-transect (the tran- sect was followed once during the day and again at night) visual sampling survey to examine differences in fish abundance, distribution, and behavior on Heceta Bank, a temperate reef and ridge ecosystem off the central coast of Oregon. Diel distributions and activity pat- terns in fishes have been well studied in tropical areas, and these studies have shown that most fishes exhibit distinct diel behavioral patterns (Col- lette and Talbot, 1972; Hobson, 1972; Helfman, 1978). Generally, two-thirds of fishes are diurnal, one-third are nocturnal, and a marked change in the vertical distribution of fishes oc- curs between day and night (Helfman, 1978). Most fishes rigidly follow a di- urnal or nocturnal activity pattern, exhibiting very low activity in shel- ter-providing areas and high activ- ity during feeding. Rotation between ecological niches at daytime and at night, such as the broad replacement of diurnal planktivores with preda- tors during the night, proceeds in predictable patterns (Hobson, 1972). Some of the largest schools encoun- tered by day in tropical waters are nonfeeding, resting schools of noctur- nal fishes, and many diurnal fishes actively school in the water column by day and then rest individually at night (Hobson and Chess, 1973; Par- rish, 1992). Although not as well studied, diel shifts in fish communities and be- havioral patterns in shallow temper- ate habitats are less distinct than Hart et al.: Abundance, distribution, and activity patterns of demersal fishes on Hecate Bank, Oregon 467 those observed in warmer water regions (Hobson et al., 1981). Deepwater habitats with low light penetra- tion would therefore be expected to show only subtle changes in species composition, densities, or activities during night hours. Nevertheless, diel distributions and activity patterns of some species of fish have been observed in temperate areas at substantial depths. On Stonewall Bank, Oregon, in situ direct observations in shelf waters (41-70 m) revealed that species composition changed little from day to night, but the abundance of some fishes decreased dramatically (Hixon and Tissot1). Specifically, juvenile rockfishes ( Sebastes spp. ) and ro- sethorn rockfishes (S. helvomaculatus) showed much greater abundance during the day, and spotted ratfish ( Hydrolagus colliei ) and widow rockfish (S. entome- las ) were significantly more abundant at night. Within Pribilof Canyon (181 to 240 m) in the Bering Sea, Pa- cific ocean perch ( S . alutus) actively fed on euphausiids just above sea-whip “forests” during the day and were observed to be less active within the sea-whip habitat at night (Brodeur, 2001). In deeper rocky bank areas with lower levels of ambient light, it is not known whether an overall change from high to low activity from day to night exists, as is observed in shallow temperate fish communities (Ebeling and Bray, 1976; Hobson and Chess, 1976; Moulton, 1977). We hypothesized there would be differences in the day and night assemblages of fishes in deep temperate waters, but that the patterns and changes in abundance and activity would be less distinct than those observed in tropical and shallow temperate fish communities. Given its diverse range of habitats and water depths, Heceta Bank is an ideal location for studying day-night patterns of demersal fishes among different depths and habitat types. Materials and methods Study area Heceta Bank is one of the largest of all submarine, rocky banks off the west coast of the United States, located approximately 60 km off the central Oregon coast, extending 50 km north to south (Fig. 1). The bank has been a primary focus of direct-observation studies of groundfishes, invertebrates, and habitat since the late 1980s (Pearcy et al., 1989; Stein et al., 1992; Wakefield et al., 2005; Whitmire et al., 2007; Hixon and Tissot1; Hixon et al.2). This bank comprises a wide range of 1 Hixon, M. A., and B. N. Tissot. 1992. Fish assemblages of rocky banks of the Pacific Northwest. Final Report supple- ment, OCS Study 91-0025, 128 p. U.S. Minerals Manage- ment Service, Camarillo, CA 93010. 2 Hixon, M. A., B. N. Tissot, and W. G. Pearcy. 1991. Fish assemblages of rocky banks of the Pacific northwest, Heceta, Coquille, and Daisy Banks. OCS Study MMS 91-0052, 410 p. U.S.D.I. Minerals Management Service, 770 Paseo Camarillo, 2nd Floor, Camarillo, CA 93010. benthic habitats from rock ridge and boulder to sand, and mud and extends from 70 m depth at the top of the bank to >500 m in water depth on its flanks (Figs. 1 and 2). The bank has been generally characterized as having three major habitat-depth profiles: 1) shallow rock ridge and boulder habitat from 70 to 100 m; 2) boulder and cobble habitat at mid-depth from 100 to 150 m; and 3) mud habitat in greater than 150 m of water depth (Hixon et al.2). Some portions of the bank show great habitat variability (Hixon and Tissot1). Light levels Ambient light levels on Heceta Bank were measured two weeks before the current study by the Plankton/ Bio-Optics group at Oregon State University during a Global Ocean Ecosystems Dynamics (GLOBEC) study of meso- and fine-scale physical and biological fields (Whit- mire and Cowles, unpubl. data3). In order to characterize the underwater light environment over the bank, the OSU group used in situ total absorption coefficient data. Absorption measurements (at[488 nm|) were collected with a dual-path absorption and attenuation meter (ac-9; WET Labs, Inc., Philomath, OR) that was mounted on a SeaSOAR, a towed undulating vehicle used to deploy a wide range of oceanographic monitoring equipment, and towed in an undulating fashion along east to west track lines over the bank during daylight hours on 5 and 6 June 2000. One percent light levels (in relation to surface light values), a commonly used oceanographic parameter for comparing light attenuation, were reached at approximately 20 m in high chlorophyll coastal waters (~3 to 6 mg/m), and at approximately 50 m in waters at the western end of the east— west transects. This empirical approach for light attenuation agreed well with theoretical relationships between depth and light attenuation as applied to coastal and offshore waters within the California Current (Morel, 1988; Barnard et al., 1999). Survey transects From 19 to 26 June 2000, an interdisciplinary group of scientists used the remotely operated vehicle (ROV) ROPOS (Remote Operated Platform for Ocean Science), managed and operated by the Canadian Scientific Sub- mersible Facility (CSSF), to revisit five stations on Heceta Bank that were established in the 1980s (Figs. 1 and 2) (Pearcy et al., 1989; Stein et al., 1992; Hixon and Tissot1; Hixon et al.2) and to explore new sites on the bank. At each of five historical stations fish assemblages were compared between day and night. The ROV ROPOS is well suited for deepwater demer- sal fish surveys. ROPOS is a 30-horsepower electro- hydraulic ROV equipped with two video systems, a 3 Whitmire, A. L., and T. J. Cowles. 2008. College of Oceanic and Atmospheric Sciences, Oregon State Univ., Corvallis, OR 97331. 468 Fishery Bulletin 108(4) 125°0'W 124°52'W 125°0'W 124°30'W 124°0'W Figure 1 Location and depth (in meters) of the study area, Heceta Bank off the Oregon coast, and the area mapped with a multibeam sonar system in 1998 (MBARI 2201; modified from Whitmire et ah, 2008). The leftmost panel shows the location of the ROV transects for areas surveyed during day and night periods for each station (boxes) on Heceta Bank, Oregon. Global Ocean Ecosystem Dynamics (GLOBEC) track lines for light levels (lines 3, 3a, 4, and 4a) are shown in white dashed lines and are labeled to the right of the map. Depth contours are given in meters. broadcast-quality Sony DXC 950 three-chip color video camera (used for fish and habitat video analysis), a wide-angle low light black and white video camera, an obstacle avoidance sonar, a compass, three arc lights (250 W), four halogen lights with adjustable intensity (250 W), and a 10-hp cage with a separate light and video system (Wallace and Shepherd, 2003). One pair of scaling lasers (10-cm scale) mounted in parallel on the color video camera provided scale in the field of view of the video image for estimating the width of transects and the size of fishes and features on the seafloor. The distance surveyed was determined from smoothed navi- gation of tracklines obtained from ultra-short baseline tracking. Real-time audio commentary of habitat type and fish identification was overlaid on the videotape. The technical side of the ROV dive transects was man- aged and conducted by two four-member CSSF teams that worked in alternating 12-hour watches. Parallel interdisciplinary science teams worked in tandem with the ROV group to direct the scientific operations and ensure consistency in sampling effort. Day and night complements were completed outside the two-hour twilight periods of dawn (0436 to 0636 PST) and dusk (2006 to 2206 PST) to avoid possible biases due to changes in the behavior of fishes dur- ing crepuscular periods (Yoklavich et al., 2000). Only daytime fish transects that overlapped geographically with, or were in close proximity to, corresponding night fish transects (stations 2, 3, 4, 6, and 9) were used in this analysis. Station 4 contained a night transect located approximately 1 nautical mile (nmi) away from the day transect but in a comparable depth range and over similar habitat, and therefore it was included in the analysis. A total area of 5.5 hectares along a combined total transect distance of 40.9 km was surveyed during 45.6 hours (Fig. 2). Stations varied in habitat composition and depth (Fig. 2). Shallow areas (70-100 m) were dominated Hart et al.: Abundance, distribution, and activity patterns of demersal fishes on Hecate Bank, Oregon 469 Station 4 Day I...,, " I "1 n R527 Night M R528 o'o 0'5 TO Station 9 Day HI R537 Night R538 0.0 0.5 TO Station 2 Day (HIT- — iuuth R531 Niaht 'I a 0.0 0.5 1.0 Station 3 Day L .1 rm R532 Night fimm R533 0.0 0.5 1.0 Station 6 Dav i -fe«.vl R534 Night L 11! Area (ha) Day Night 11 Flat Rock B Flat Rock 13 Rock ridge □ Rock ridge E] Boulder H Boulder □ Cobble H Cobble D Pebble 02 Pebble □ Sand D Sand □ Mud □ Mud Figure 2 Bar graphs indicate the area and relative composition (in hectares) of each primary habitat surveyed along transects during day and night periods for each station on Heceta Bank, OR. ROV dives associated with each station are identified by an “R” followed by the dive number. by rock ridge and boulder, mid-depth by cobble, and deeper areas by mud habitat with isolated patches of cobble and boulder. Rock ridge, boulder, cobble, and mud composed the four most dominant primary habitat types. Videotape analysis Videotape collected along all transects was analyzed for fish and habitat identification. Sunset, dawn, dusk, and nautical twilight times (Pacific daylight savings time) were derived from the U.S. Naval Observatory website (http://www.usno.navy.mil/, accessed June 2000) and calculated for the day of each dive by using the specific longitude and latitude coordinates (degrees and minutes) for each station. Transects were subdivided into habitat patches, based on primary habitat types observed on the videotapes. Seafloor habitats were classified into seven standard- ized categories (Hixon and Tissot1; Stein et al., 1992): rock ridge (high relief where vertical rock was found to be >3.0 m); flat rock (low relief where vertical rock was found to be <3.0 m); boulder (300-25.6 cm rock); cobble (25.6-6.4 cm rock); pebble (6. 4-0. 2 cm rock); sand (2.0-0.06 mm); and mud (<0.06 mm). Only the pri- mary habitat (>50% of the seafloor) in the field of view was used in our analysis. The length of each habitat patch was determined by using the geographic position recorded at the start and end of each patch. With the scaling lasers, the width of each transect was estimated by selecting random frames every minute during each transect, measuring the width of these lasers on the video monitor, and extrapolating to the field of view. Transect width ranged from 1.3 to 2.1 meters. The area of each patch was determined by multiplying the patch length by the average patch width. Each transect consisted of a few to many habitat patches, depending on the variability of substrate. Videotape analysis was performed by two technicians simultaneously in order to confirm fish identifications, counts, and fish activity. All fishes were identified to the lowest practical taxonomic unit (usually to species) and total fish length was estimated to the nearest 5 cm. For fish that could be identified to a taxonomic group, but not species, a generalized abbreviation was used (e.g., FF for unidentified flatfish), and in cases where the fish observed was one of two species, a new abbreviation was created to accommodate this situa- tion. Fish were counted at the point where they passed through the level of the scaling lasers and were as- signed a GMT time that became a permanent time and geographic reference point in a database. A single ana- tomical feature (eye) was used to determine whether or not a fish was considered within the transect to prevent underestimating transect width and overestimating abundance. We restricted our analysis to 31 identified species or taxonomic groups that were seen frequently enough to represent >0.1% of the total day and night fish density (number of fish per hectare). Exceptions to this rule were the inclusion of three rarely seen species: darkblotched rockfish (S. crameri), because this is an important commercial species; kelp greenling (Hexa- grammos decagrammus ); and bigfin eelpout ( Lycodes cortezianus), because the day-night activity patterns of the latter two have been studied in shallow temperate areas (Moulton, 1977). Some categories of multiple taxa were created, such as “pygmy-Puget Sound rockfish complex” {S. wilsoni and S. emphaeus), where it was impossible to identify every individual in aggregations. Another common category was “unidentified rockfish,” where it was not possible to classify a rockfish to spe- cies conclusively. “Unidentified juvenile rockfish” (ab- breviation RRF) were categorized by size (<10 cm long), not by morphological features, because video resolu- tion and inherent difficulty of in situ identification of young-of-the-year rockfishes precludes determinations at the level of species. Hagfishes (Eptatretus spp. ) were not identified to species, but on the basis of depth of 470 Fishery Bulletin 108(4) occurrence, the species observed was probably the Pa- cific hagfish (Eptatretus stoutii) (Barss, 1993). All fish in the videotapes were counted and assigned an activity category (“active” or “inactive”). The two cat- egories were developed in order to analyze the videotape efficiently and quantitatively, and all fishes were placed into one of the two categories. Active fish were off the bottom (or temporarily in contact with the substrate), and inactive fish were in contact with the substrate (e.g., in contact with the seafloor or were occupying crevices). All flatfish were excluded from the activity analysis because of our definition of “activity.” We as- sumed that fish were counted only once and that the submersible did not influence the activity of fishes. When the ROV clearly affected the behavior of a fish, the activity of the fish observed before the ROV inter- ruption was used in analyses. Data analysis Relative abundance was determined for each taxon on a broad scale over all stations (2, 3, 4, 6, and 9) and all habitat types, and on a finer scale within each of the four primary habitat types (rock ridge, boulder, cobble, and mud) over all stations. Fish abundance was first normalized by dividing the abundance for a given taxon in a habitat patch by the area swept in that habitat patch. Relative abundance in a daytime habitat patch was matched up with a relative abundance in a night- time habitat patch of closest geographic proximity for a given taxon. These pairs were used to create the ranks in a Wilcoxon signed-rank test. For all taxa, 402 habitat patches were included in the analysis, and an average of 59 habitat patches were compared for each taxon (many taxa were in the same habitat patch). This enabled us to estimate if relative abundance trends were consistent at both scales. For the finer scale, some habitat types (e.g., flat rock) did not contain sufficient relative abundance for analysis. The Wilcoxon signed-rank test was performed with S-Plus, vers. 3.2 software (TIBCO Software Inc., Palo Alto, CA) to determine significant differences in the ac- tivity of fishes during day and night (Ramsey and Scha- fer, 2002). This test involved estimating the percentage of fish active and inactive within each habitat patch for each taxon (raw abundance was used). These day and night pairs (percentage of fish active and percentage of fish inactive) were used to create the ranks in the test for a given taxon over primary habitat types. A total of 398 habitat patches were analyzed for all taxa, with an average of 64 habitat patches compared for each taxon. We also used nonmetric multidimensional scaling (NMS) to examine associations between day and night fish abundance with depth and with primary habitat (McCune and Grace, 2002). PC-ORD software, vers. 5.0 (MjM Software, Gleneden Beach, OR), was used with a Monte Carlo test and Sprensen distance mea- sure, starting with random configurations (Mather, 1976). We restricted the NMS to taxa that showed significantly greater abundance during day or night in the Wilcoxon signed-rank test (P<0.05), and to those that showed a strong correlation (Pearson and Kendall correlation) with depth (P<-0.5 or >0.5 on the second axis) during trial NMS runs with all 31 taxa. A total of 11 taxa met these criteria. The final species (taxa) matrix included columns of log-transformed abundance and rows of sample units grouped by primary habitat for each dive during day and night (e.g., boulder-night - R534). All primary habitat types were used because the NMS determines correlation strength along an environmental gradient and does not require paired plots as in the Wilcoxon signed-rank test. The final environmental matrix included two quantitative vari- able columns: primary habitat types (flat rock, rock ridge, boulder, cobble, pebble, sand, and mud) and average depth (meters). Sample units greater than 3.0 standard deviations were excluded from analysis. The final ordination had 166 iterations and 15 runs. This test enabled us to determine if marked differences in fish abundance were associated with depth and pri- mary habitat, and whether taxa showing significantly greater abundance during day or night were distrib- uted similarly on the bank. Results A total of 29,787 individual fish were counted on the ROV transect videotapes. During the day, we observed an average of 207 fishes per hectare, and at night we observed a lower average of 141 fishes per hectare. Fish taxa in greatest abundance were from the genus Sebastes. Dominant taxa (pygmy and Puget Sound rock- fish, and unidentified juvenile rockfish) showed the largest differences in relative abundance between day and night (Fig. 3). Across all stations and primary habitat types, and within at least one primary habitat type, eight taxa showed significantly greater abundance during the day (P<0.05) and five taxa exhibited sig- nificantly greater abundance during the night (P<0.05, Table 1 , Fig. 3). Three taxa were found to be signifi- cantly greater in abundance during day (kelp greenling and unidentified mottled poacher [Agonidae]) and night (redstripe rockfish [S. proriger ] ) only in specific primary habitat types (Table 1). Several taxa showed apparent differences in abundance, but sample sizes were too small for statistical significance in the paired Wilcoxon signed rank test (e.g., kelp greenling). Harlequin rock- fish ( S . variegatus) was the only species we regularly encountered exclusively at night (darkblotched rockfish were rare but were also seen only at night), whereas kelp greenling were encountered only during the day at shallow depths. NMS analysis of distribution The NMS analysis showed significant correlations (P=0.03) among taxa, depth, primary habitat, and day and night (Fig. 4). The ordination explained 78% of the variation with an acceptable stress value (a lower value Hart et al.: Abundance, distribution, and activity patterns of demersal fishes on Hecate Bank, Oregon 471 "Pygmy rockfish [ "Puget Sound rockfish [ ‘Pygmy-Puget Sound complex [ "Unident, juvenile rockfish [ "Rosethorn rockfish [ "Sharpchin rockfish [ Unident adult rockfish [ Yellowtail rockfish f Greenstriped rockfish j Redstripe rockfish [ "Harlequin rockfish Widow rockfish Yelloweye rockfish Canary rockfish Darkblotched rockfish Shortspine thornyhead || ~ | , Dover sole Unident, flatfish ‘Rex sole "Unident, sculpin "Unident, mottled sculpin Threadfin sculpin Icelinus spp Kelp greenling Lingcod ‘Eptatretus sp. "Spotted ratfish Sablefish Unident, mottled poacher ‘Unident, ronquil br.--”r'T Bigfin eelpout | 1 33 =11 Ha 11 3*111 3* □ Day O Night :a±- IV =Bh 10 100 Number of fish/hectare 1000 10,000 Figure 3 Total abundance (number of fish per hectare) of fish taxa as determined from the ROV during night and daytime transects over all habitat types. Wilcoxon signed-rank test was used to compare day and night abundance across stations and primary habitat types (*P<0.05, ** PcO.Ol). Arabic numerals indicate the most dominant taxa during day, and Roman numer- als indicate the most dominant taxa during night. Error bars indicate standard error of the abundance for each taxon. White bars represent day abundance and gray bars represent night abundance for the following species groups: Sebastes, Sebastolobus, flatfishes, Cottidae, and other fish. Scientific names for all taxa include from top to bottom: pygmy rockfish (S. wilsoni ), Puget Sound rockfish (S. emphaeus ), unidentified juvenile rockfish less than 10 cm long (Sebastes spp.), rosethorn rockfish (S. helvomaculatus), sharpchin rockfish (S. zacentrus), unidentified adult rockfish (Sebastes spp.), yellowtail rockfish (S. flavidus), greenstriped rockfish (S. elongatus), redstripe rockfish (S. proriger), harlequin rockfish (S. variegatus), widow rockfish (S. entomelas ), yelloweye rockfish (S. ruberrimus), canary rockfish (S. pin- niger ), darkblotched rockfish (S. crameri), shortspine thornyhead (S. alascanus ), Dover sole (Microstomus pacificus), unidentified flatfish (Pleuronectidae), rex sole (Glyptocephalus zaclii- rus ), unidentified sculpin (Cottidae), unidentified mottled sculpin (Cottidae), threadfin sculpin (Icelinus filamentosus), Icelinus spp., kelp greenling (Hexagrammos decagrammus ), lingcod ( Ophiodon elongatus), unidentified hagfishes (Eptatretus spp.), spotted ratfish (Hydrolagus colliei), sablefish (Anoplopoma fimbria), unidentified mottled poacher (Agonidae), unidentified ronquil (Bathymasteridae), and bigfin eelpout (Lycodes cortezianus) . Alternate shading of the background represents general taxonomic groups. 472 Fishery Bulletin 108(4) Table t Fish taxa exhibiting significantly greater relative abundance (number of fish per hectare) across all stations over primary habi- tat types at Heceta Bank, OR, (rock ridge, boulder, cobble, and mud) during daytime or nighttime. Wilcoxon signed-rank test was used (*P<0.05, **P<0.01) to compare day and night relative abundance between habitat patches in closest geographic proximity within primary habitat types over all stations. Taxa are listed in order of relative abundance from top to bottom; primary habitat is listed in decreasing order of size from left to right, and numbers in parentheses are relative abundance during day or night. Taxon Rock ridge Boulder Cobble Mud Significantly more abundant during day unidentified juvenile rockfish ( Sebastes spp.) (1498)** (2123)* (632)** (150)* Puget Sound rockfish (S. emphaeus ) (2235)** (421)** pygmy rockfish (S. wilsoni ) (4878)** (46)* (634)** pygmy-Puget Sound complex (S. wilsoni and S', emphaeus ) (3240)* rosethorn rockfish (S. helvomaculatus) (249)** (291)* (124)* unidentified ronquil (Bathymasteridae) (16)** (43)* unidentified mottled sculpin (Cottidae) (30)** unidentified sculpin (Cottidae) (12)* (28)** unidentified mottled poacher (Agonidae) (12)* kelp greenling ( Hexagrammos decagrammus ) (6)* Significantly more abundant during night sharpchin rockfish (S. zacentrus ) (460)** (1134)* (5603)* (1656)* redstripe rockfish (S. proriger ) (90)* harlequin rockfish (S. variegatus) (103)** spotted ratfish (Hydrolagus colliei) (75)* (56)* unidentified hagfishes ( Eptatretus spp.) (45)* (64)* rex sole (Glyptocephalus zachirus) (69)* denotes a better “fit” of data) of 20.5 for three axes. After a +115° rotation, the first axis showed correlation with day (right side) and night (left side) sample units, explaining 22% of the variation in the data and revealed a mean stress of 49.8. The second axis explained an additional 30% of the variation in the data and showed a negative correlation with depth (Pearson’s r=0.141) and a mean stress of 29.1. The second axis also showed a positive correlation with substrate, and larger-size primary habitat was found in the first and second quad- rants of the graph and smaller-size primary habitat in the third and fourth quadrant. The third axis improved the cumulative coefficient of determination, r2, to 0.777 with little additional stress; this axis enabled us to rotate the ordination in three dimensions to distinguish habitat and species and showed good correlations with the day-night and habitat-depth axes, but is not easily plotted. Higher dimensions showed little improvement in model fit. Taxa more abundant during the day (rosethorn rock- fish, pygmy rockfish, pygmy-Puget Sound rockfish com- plex, kelp greenling, and unidentified juvenile rockfish) showed a positive correlation along axes one and two and appear in the upper right quadrant of Figure 4. Puget Sound rockfish showed a correlation with the day-night axis, but this species was associated with greater depth and smaller-size substrate as primary habitat when compared to the other taxa that were more abundant during the day; this taxon appears in the lower right quadrant. Thus, this dominant day assemblage was observed mainly at shallow- to mid- depths over medium to large-size substrata. Taxa show- ing greater abundance during the night (spotted ratfish [Hydrolagus colliei], hagfishes, rex sole [Glyptocephalus zachirus], sharpchin rockfish [S. zacentrus], and harle- quin rockfish) showed a negative correlation along axes one and two. Thus, the dominant night assemblage was generally over deeper areas of medium- to small-size substrata of cobble and mud. Day and night species assemblages During the day within all stations, large densities of mostly small-size rockfish taxa were primarily found over shallow rock ridge, boulder, and cobble substrata (Table 1, Figs. 3 and 4). The four most dominant day taxa (both active and inactive) were pygmy rockfish, Puget Sound rockfish, pygmy-Puget Sound rockfish complex, and unidentified juvenile rockfish. Yellowtail rockfish (S. flavidus) were also an important component of the daytime assemblage, albeit less abundant and less dominant during night. Many of these taxa showed sig- nificantly greater abundance and activity over medium- to large-size habitat (cobble, boulder, and rock ridge) (Tables 1 and 2). Rosethorn rockfish showed the clearest diurnal pattern of all species (Table 2). Small rockfishes and yellowtail rockfish were more active during the day but did not exhibit clear inactivity during the night. Hart et al.: Abundance, distribution, and activity patterns of demersal fishes on Hecate Bank, Oregon 473 Larger A CD n C/5 .2 ro to -Q D co A Day— Y Night X Day- + Night O Day— ♦ Night O Day—' • Night □ Day— a Night A Day— A Night x Day- mud —mud sand — sand pebble pebble cobble —cobble boulder —boulder rock ridge rock ridge flat rock ▼ Smaller Night <- ' » PR ° A s JV* A “> A "*• ® „ • RA • ;* \HF R$\ 9 RT o ° ° CKG , & ■ % A A SH HR S 4. A o c PRC RRF PSR O Shallow A Axis 1 T Deep -> Day Figure 4 Nonmetric multidimensional scaling (NMS) ordination showing association between 11 taxa and habitat patches (different symbols) from all stations (2, 3, 4, 6, and 9). Taxa codes are: HF (hagfishes, Eptatretus spp.), HR (harlequin rockfish, S’. variegatus), KG (kelp greenling, Hexagrammos decagrammus), PR (pygmy rockfish, S. wilsoni), PRC (pygmy-Puget Sound complex, S’. wilsoni and S. emphaeus), PSR (Puget Sound rockfish, S. empha- eus), RA (spotted ratfish, Hydrolagus colliei), RS (rex sole, Glyptocephalus zachirus), RT (rosethorn rockfish, S. helvomaculatus), RRF (unidentified juvenile rockfish, Sebastes spp.), and SH (sharpchin rockfish, S. zacentrus ). Environmental variables include the quantitative variable of depth (increasing depth from top to bottom), and the categorical variables (shape of symbols) of day, night, and habitat. When compared to the dominant day assemblage, the night assemblage exhibited lower overall densities and consisted of less active, larger-size fish of fewer taxa, many of which were also seen during the daylight hours (Tables 1 and 2, Fig. 3). Dominant night taxa were sharpchin rockfish, yellowtail rockfish, Dover sole (Microstomus pacificus), and greenstriped rockfish (S. elongatus). Of the dominant night assemblage, sharp- chin rockfish comprised over half of the fish encoun- tered at night (Fig. 3) and was one of the only taxa that showed significantly greater abundance during the night (P<0.05) in most primary habitat types (Table 1). Hagfishes and spotted ratfish were significantly greater in abundance and activity during the night (P<0.05), and widow rockfish showed significantly greater activity during the night (P<0.05), despite being rarely observed (Table 2). Hagfishes were the only taxa that exhibited a distinct nocturnal activity pattern (P<0.05). Discussion The relative composition of fish taxa over all depths and habitat types on Heceta Bank did not show a broad replacement from day to night, as observed in shallow tropical areas. However, there were consistent patterns of abundance and activity that were species-, depth-, and habitat-specific. There was a considerable day- night change in the abundance of the four most domi- nant day taxa (pygmy rockfish, Puget Sound rockfish, pygmy-Puget Sound rockfish complex, and unidentified juvenile rockfish) over shallower areas of rock ridge and boulder habitat. Active, small- to medium-size fish taxa tended to aggregate around shallow medium- to large-size habitat features during the day, and larger- size night taxa (sharpchin rockfish, yellowtail rockfish, Dover sole, and greenstriped rockfish) tended to aggre- gate around these features at night. Rosethorn rockfish 474 Fishery Bulletin 108(4) Table 2 Fish taxa (raw count s) exhibiting a significant difference in the percentage of fish found to be active and inactive during day and night within similar primary habitat types (rock ridge, boulder, cobble, and mud) over all stations at Heceta Bank, OR. Values represent the percentage of fish found to be active or inactive over each substrate type, and values in parentheses indicate n for each comparison. Wilcoxon signed-rank test was used (*P<0.05, **P<0.01, not significant [ns], P>0.05). Taxa are listed in order of highest abundance (number of fish per hectare) within each category. Taxon Primary habitat Day active fish Day inactive fish Night active fish Night inactive fish Diurnal rosethorn rockfish (S. helvomaculatus) rock ridge 76% (246)** ns boulder 74% (243)** 85% (123)** cobble 66% (405)* 92% (143)** mud ns 93% (47)** Nocturnal Eptatretus spp. cobble 87% (23)* 69% (29)* Significantly more active during day unidentified juvenile rockfish rock ridge 99% (855)** ( Sebastes spp.) boulder 98% (1214)** cobble 96% (876)** mud 91% (54)** Puget Sound rockfish (S. emphaeus) boulder 71% (367)** cobble 73% (2501)** mud 87% (126)** pygmy rockfish (S. wilsotii) rock ridge 99% (855)** cobble 99% (235)** mud 94% (34)** yellowtail rockfish (S. flavidus) rock ridge 95% (259)** boulder 96% (260)** Significantly more active during night spotted ratfish (Hydrolagus colliei ) boulder 87% (16)* mud 92% (87)* widow rockfish (S. entomelas) boulder 96% (45)* and hagfishes exhibited distinct diurnal and nocturnal activity, respectively. These day and night patterns were similar to those observed during day-time surveys from manned submersibles on Heceta bank (Pearcy et al., 1989; Stein et al., 1992; Hixon et al.2), day and night surveys on Stonewall Bank (Hixon and Tissot1), and are generally consistent with most patterns found in other shallow temperate day and night studies, but were much less distinct than those for fishes inhabiting tropical fish communities (Helfman, 1978). The overall marked decrease in abundance and activity of smaller-size taxa at night was similar to the decrease that Ebeling and Bray (1976) and Moulton (1977) observed, but our study did not provide evidence of a pronounced replacement of diurnally active taxa by exclusively nocturnal spe- cies as observed at Santa Catalina Island (Hobson and Chess, 1973; Hobson et al., 1981), Hawaiian tropical reefs (Hobson, 1972), and reefs in the Virgin Islands (Colette and Talbot, 1972). It is possible that light illumination at the top of Heceta Bank during the day contributed to the higher abundance and activity of the three most dominant day taxa (pygmy rockfish, pygmy-Puget Sound rockfish complex, and unidentified juvenile rockfish), as found in similar studies on temperate species. Fishes found at the top of Heceta Bank likely perceive and use the faint sun illumination during the day (Boehlert, 1979). On Heceta Bank, the photic zone generally extends down to approximately 50 m water depth and it is generally accepted that sun illumination affects behavior of fishes down to these depths (L. Britt, personal commun.4). This was confirmed by the GLOBEC survey, which mea- sured one percent of surface light at 50 m, just above the shallowest fish survey depths (70 m) where most of the unidentified juvenile rockfish were present. Light il- lumination may be aiding the dominant day assemblage because these taxa stay close (perhaps within visual distance) to large features that provide refuge from larger, active piscivorous predators. In Puget Sound 4 Britt, Lyle L. 2007. Alaska Fisheries Science Center, National Marine Fisheries Service, National Oceanic and Atmospheric Administration, 7600 Sand Point Way N.E., Seattle, WA 98115. Hart et al.: Abundance, distribution, and activity patterns of demersal fishes on Hecate Bank, Oregon 475 (Moulton, 1977), on rocky kelp forests off Santa Bar- bara, California (Ebeling and Bray, 1976), and on Santa Catalina Island, California (Hobson and Chess, 1976; Hobson et al., 1981), the small-size adult and juvenile rockfish that remain exposed from refuge during the night stay close to the seafloor. Only two taxa showed distinct day and night activity patterns (diurnal: rosethorn rockfish and nocturnal: hagfish), indicating that marked differences in day and night activity are not as prevalent as those found in shallow tropical coral reef and temperate fish com- munities (Table 2) (Hobson, 1972; Helfman, 1978). The diurnal activity pattern exhibited by rosethorn rockfish was independent of habitat, indicating that habitat type may not be a significant factor in determining differ- ences between day and night activities for this species (Tables 1 and 2). The nocturnal activity pattern that hagfishes exhibited was similar to that observed by Ooka-Souda et al. (1985) of another hagfish species (Eptatretus burgeri ), in the laboratory. During the day on Heceta Bank, most hagfish were observed coiled up over mud or sand, or around cobble, whereas during the night fish of this taxon was observed swimming above bottom or moving in contact with the bottom in a twisted manner. Like silver hake ( Merluccius bi- linearis) which use sand waves (transverse ridges of sand) for refuge during the day (Auster et al., 2003) and that forage during the night (Bowman and Bow- man, 1980; Auster et al., 1995), hagfishes on Heceta Bank may use mid-depth cobble and mud habitats for resting during the day and may forage at night. Not all rosethorn rockfish and hagfishes strictly followed a diurnal or nocturnal activity pattern; however, the fish activity measure we used was not sensitive enough to detect subtle differences in behavior, such as resting individuals found hovering close to but not in contact with the seafloor. Potential bias exists in observing and attempting to quantify activity in fishes when using video survey methods (Uzmann et al., 1977; Sale and Douglas, 1981; Wakefield and Smith, 1990). Our inability to identify many of the rockfishes to species is potentially problem- atic, and illustrates a limitation of this survey method. The majority of historical studies show that very few fishes exhibit changes in activity with the presence of an ROV or submersible, although a handful of taxa do show behavioral responses (High, 1980; Carlson and Straty, 1981; Pearcy et al., 1989). SCUBA-based video surveys of fish abundance and behavior indicate that although the majority of fishes show no noticeable reac- tion, some species may avoid or be attracted to divers outside the camera’s field of visibility and may even follow divers (Moulton, 1977). It has been argued that SCUBA surveys do not significantly affect counts be- cause most fishes that follow divers remain behind the field of view of the camera (Powles and Barans, 1980). In our study, anecdotal evidence indicated that the ROV had limited effect on fish behavior, except in cases where the ROV came in contact with the substrata. Further investigation is needed, however, to fully grasp the impacts of observational vehicles on fish responses (Stoner et al., 2008). Implications for groundfish surveys Day and night activity patterns in demersal fishes have been shown to dramatically change the catchability of some species on the West Coast (Hannah et al., 2005), in the Northwest Atlantic (Bowman and Bowman, 1980), in Newfoundland (Casey and Myers, 1998), and in the North Sea (Petrakis et al., 2001). In this study, we found that daytime surveys could underestimate the abundance of certain species that are more abundant or active at night, such as sharpchin rockfish. Highly significant differences in day and night abundance of schooling rockfishes found in this study indicate that daytime trawl surveys over small- to medium-size habitat features may be biased for some fish species. Migration of fishes into the overlying water column, hori- zontally off the bank, or into hiding among medium- to large-size features is likely the most common day and night behaviors that would decrease the availability of fishes to the ROV. Specifically, Puget Sound rockfish may be more available over deeper, smaller-size rock struc- tures, whereas other dominant day taxa (pygmy rockfish, pygmy-Puget Sound rockfish complex, and unidentified juvenile rockfish) and sharpchin rockfish are likely less available to trawl surveys over large-size features in shallower portions of the bank. We speculate that the reduction in abundance of the four most dominant day taxa is due to fish seeking refuge around medium- to larger-size structures because of the potential presence of large piscivores (Wilkins, 1986; Adams, 1987), rather than to schooling in the water column. Acknowledgments We would especially like to thank the following col- leagues who contributed to this study: B. Barss, B. Embley, G. Hendler, M. Hixon, D. Markle, B. McCune, S. Merle, B. Tissot, M. Yoklavich, and K. York. A. Whitmire and T. Cowles generously contributed information on ambient light levels. L. Britt, C. Whitmire, and L. Cian- nelli provided constructive reviews of the manuscript. This portion of the Heceta Bank project was funded by the West Coast and Polar Regions Undersea Research Center of the National Oceanographic and Atmospheric Administration’s (NOAA) National Undersea Research Program, the Northwest and Southwest Fisheries Sci- ence Centers, NOAA’s Pacific Marine Environmental Laboratory, and the Cooperative Institute for Marine Resources Studies at Oregon State University. We would like to thank the professional personnel who operated the ROV ROPOS and the NOAA RV Ronald Brown. T. Hart was supported through the Oregon Agricultural Experiment Station project ORE00102, the H. Richard Carlson Scholarship, the Collaborative Marine Fisher- ies Fellowship and the Bill Wick Award through Oregon State University. 476 Fishery Bulletin 108(4) Literature cited Adams, P, B. 1987. The diet of widow rockfish Sebastes entomelas in northern California. In Widow rockfish: proceedings of a workshop, Tiburon, California, December 11-12, 1980 (W. H. Lenarz, and D. R. Gunderson, eds.), p. 37-41. U.S. Dep. Commer., NOAA Tech. Rep. NMFS-48. Adams, P. B., J. L. Butler, C. H. Baxter, T. E. Laidig, K. A. Dahlin, and W. W. Wakefield. 1995. Population estimates of Pacific coast groundfish from video transects and swept-area trawls. Fish. Bull. 93:446-455. Auster, P. J., J. Lindholm, S. Schaub, G. Funnell, L. S. Kaufman, and P. C. Valentine. 2003. Use of the sand wave habitats by silver hake. J. Fish Biol. 62:143-152. Auster, P. J., R. J. Malatesta, and S. C. LaRosa. 1995. Patterns of microhabitat utilization by mobile mega- fauna on the southern New England (USA) continental shelf and slope. Mar. Ecol. Prog. Ser. 127:77-85. Barnard, A. H., J. R. V. Zaneveld, W. S. Pegau, J. L. Mueller, H. Maske, R. Lara-Lara, S. Alvarez Borrego, R. Cervantes- Duarte, and E. Valdez-Holquin. 1999. The determination of PAR levels from absorp- tion coefficient profiles at 490 nm. Ciencias Marinas 25:487-507. Barss, W. H. 1993. Pacific hagfish, Eptatretus stouti, and black hag- fish, E. deani : The Oregon fishery and port sampling observations. Mar. Fish. Rev. 55:19-30. Boehlert, G. W. 1979. Retinal development in postlarval through juvenile Sebastes diploproa : adaptations to a changing photic environment. Rev. Can. Biol. 38(4):265-280. Bowman, R. E., and E. W. Bowman 1980. Diurnal variation in the feeding intensity and catch- ability of silver hake ( Merluccius bilinearis). Can. J. Fish. Aquat. Sci. 37:1565-1572. Brodeur, R. D. 2001. Habitat-specific distribution of Pacific ocean perch (Sebastes alutus) in Pribilof Canyon, Bering Sea. Cont. Shelf Res. 21:207-224. Carlson, H. R., and R. R. Straty. 1981. Habitat and nursery grounds of Pacific rockfish, Sebastes spp., in rocky coastal areas of Southeastern Alaska. Mar. Fish. Rev. 43:13-19. Casey, J. M., and R. A. Myers. 1998. Diel variation in trawl catchability: is it as clear as day and night? Can. J. Fish. Aquat. Sci. 55:2329-2340. Collette, B. B., and F. H. Talbot. 1972. Activity patterns of coral reef fishes with emphasis on nocturnal-diurnal changeover. In Results of the Tektite program: Ecology of coral reef fishes (B. B. Collette and S. A. Earle, eds.), p. 98-124. Bull. Los Ang. Cty. Nat. Hist. Mus. Sci. 14. Ebeling, A. W., and R. N. Bray. 1976. Day versus night activity of reef fishes in a kelp forest off Santa Barbara, California. Fish. Bull. 74:703-717. Gunderson, D. R. 1993. Surveys of fisheries resources. 256 p. John Wiley and Sons, New York. Hannah, R. W., S. J. Parker, and T. V. Buell. 2005. Evaluation of a selective flatfish trawl and diel variation in rockfish catchability as bycatch reduction tools in the deepwater complex fishery off the U.S. west coast. N. Am. J. Fish Manag. 25:581-593. Helfman, G. S. 1978. Patterns of community structure in fishes: sum- mary and overview. Environ. Biol. Fishes 3:129-148. High, W. L. 1980. Bait loss from halibut longline gear observed from a submersible. Mar. Fish. Rev. 42(2):26-29. Hobson, E. S. 1972. Activity of Hawaiian reef fishes during the evening and morning transitions between daylight and darkness. Fish. Bull. 70:715-740. Hobson, E. S, and J. R. Chess. 1973. Feeding oriented movements of the Athernid fish Pranesus pingus at Majoru atoll, Marshall Islands Fish. Bull. 71:777-786. 1976. Trophic interactions among fishes and zooplank- ters nearshore at Santa Catalina Island. Cal. Fish. Bull. 74:567-598. Hobson, E. S., W. N. McFarland, and J. R. Chess. 1981. Crepuscular and nocturnal activities of California nearshore fishes, with consideration of their scotopic visual pigments and the photic environment. Fish. Bull. 79:1-30. Mather, P. M. 1976. Computational methods of multivariate analy- sis in physical geography, 532 p. J. Wiley and Sons, London. McCune, B., and J. B. Grace. 2002. Analysis of ecological communities, 304 p. MjM Software, Gleneden, OR. Morel, A. 1988. Optical modeling of the upper ocean in relation to its biogenous matter content (case I waters). J. Geo- phys. Res. 93:10749-10768. Moulton, L. L. 1977. An ecological analysis of fishes inhabiting the rocky nearshore regions of northern Puget Sound, Washington. Ph.D. diss., 181 p. Univ. Washington, Seattle, WA. Ooka-Souda, S., H. Kabasawa, and S. Kinoshita. 1985. Circadian rhythms in locomotor activity in hag- fish, Eptatretus burgeri, and the effect of reversal of light-dark cycle. Zool. Sci. 2:749-754. Parrish, J. K. 1992. Levels of diurnal predation on a school of flat-iron herring, Harengula thrissina. Environ. Biol. Fishes 34:257-263. Petrakis, G., D. N. MacLennan, and A. W. Newton. 2001. Day-night and depth effects on catch rates during trawl surveys in the North Sea. ICES J. Mar. Sci. 58:50-60. Pearcy, W. G., D. L. Stein, M. A. Hixon, E. K. Pikitch, W. H. Barss, and R. M. Starr. 1989. Submersible observations of deep-reef fishes of Heceta Bank, Oregon. Fish. Bull. 87:955—965. Powles, H., and C. A. Barans. 1980. Groundfish monitoring in sponge-coral areas off the southeastern United States. Mar. Fish. Rev. 42 ( 5 ) : 2 1— 35 . Ramsey, F., and D. Schafer. 2002. The statistical sleuth: a course in methods of data analysis, 2nd ed., 816 p. Duxbury Press, Pacific Grove, CA. Sale, P. F., and W. A. Douglas. 1981. Precision and accuracy of visual census technique Hart et al.: Abundance, distribution, and activity patterns of demersal fishes on Hecate Bank, Oregon 477 for fish assemblages on coral patch reefs. Environ. Biol. Fishes 6:333-339. Stein, D. L., B. N. Tissot, M. A. Hixon, and W. Barss. 1992. Fish-habitat associations on a deep reef at the edge of the Oregon continental shelf. Fish. Bull. 90:540-551. Stoner, A. W., C. H. Ryer, S. J. Parker, P. J. Auster, W. W. Wakefield. 2008. Evaluating the role of fish behavior in surveys conducted with underwater vehicles. Can. J. Fish. Aquat. Sci. 65:1230-1243. Uzmann, J. R., R. A. Cooper, R. B. Theroux, and R. L. Wigley. 1977. Synoptic comparison of three sampling techniques for estimating abundance and distribution of selected megafauna: submersible vs. camera sled vs. otter trawl. Mar. Fish. Rev. 39:11-19. Wakefield, W. W„ and K. L. Smith Jr. 1990. Ontogenetic vertical migration in Sebastolobus altivelis as a mechanism for transport of particulate organic matter at continental slope depths. Limnol. Oceanogr. 35:1314-1328. Wakefield, W. W., C. E. Whitmire, J. E. R. Clemons, and B. N. Tissot. 2005. Fish habitat studies: combining high-resolution geo- logical and biological data. In Benthic habitats and the effects of fishing (P. W. Barnes and J. P. Thomas, eds.), p. 119-138. Am. Fish. Soc. Symp. 41, Bethesda, MD. Wallace, K., and K. Shepherd. 2003. The streaming data management challenge: inte- grating multiple channels of real-time data from various sources in a flexible and familiar environment. Sea Tech. 44:37-44. Wilkins, M. E. 1986. Development and evaluation of methodologies for assessing and monitoring the abundance of widow rockfish, Sebastes entomelas. Fish. Bull. 84:287—310. Whitmire, C. E., R. W. Embley, W. W. Wakefield, S. G. Merle, and B. N. Tissot 2007. A quantitative approach for using multibeam sonar data to map benthic habitats. In Mapping the seafloor for habitat characterization (B. J. Todd and H. G., Greene, eds.), p. 111-126. Geol. Assoc. Can. Spec. Pap. 47. Yoklavich, M. M., H. G. Greene, G. M. Cailliet, D. E. Sullivan, R. N. Lea, and M. S. Love. 2000. Habitat associations of deep-water rockfishes in a submarine canyon: an example of a natural refuge. Fish. Bull. 98:625-641. 478 Abstract — Ichthyoplankton surveys were conducted in shelf and slope waters of the northern Gulf of Mexico during the months of May— September in 2005 and 2006 to investigate the potential role of this region as spawn- ing and nursery habitat of sailfish ( Istiophorus platypterus). During the two-year study, 2426 sailfish larvae were collected, ranging in size from 2.0 to 24.3 mm standard length. Mean density for all neuston net collections (n= 288) combined was 1.5 sailfish per 1000 m2, and maximum density was observed within frontal features cre- ated by hydrodynamic convergence (2.3 sailfish per 1000 m2). Sagittal otoliths were extracted from 1330 larvae, and otolith microstructure analysis indicated that the sailfish ranged in age from 4 to 24 days after hatching (mean=10.5 d, standard de- viation [SD] = 3.2 d). Instantaneous growth coefficients ( g ) among survey periods (n = 5) ranged from 0.113 to 0.127, and growth peaked during July 2005 collections when density within frontal features was high- est. Daily instantaneous mortality rates ( Z ) ranged from 0.228 to 0.381, and Z was indexed to instantaneous weight-specific growth (G) to assess stage-specific production potential of larval cohorts. Ratios of G to Z were greater than 1.0 for all but one cohort examined, indicating that cohorts were gaining biomass during the majority of months investigated. Stage-specific production potential, in combination with catch rates and densities of larvae, indicates that the Gulf of Mexico likely represents important spawning and nursery habitat for sailfish. Manuscript submitted 4 March 2010. Manuscript accepted 28 July 2010. Fish. Bull. 108:478-490 (2010). The views and opinions expressed or implied in this article are those of the author (or authors) and do not necessarily reflect the position of the National Marine Fisheries Service, NOAA. Distribution, growth, and mortality of sailfish (Istiophorus platypterus ) larvae in the northern Gulf of Mexico Jeffrey R. Simms1 (contact author) Jay R. Rooker1 Scott A. Holt2 G. Joan Holt2 Jessica Bangma3 Email address for contact author: jsimms@entrix.com 1 Department of Marine Biology Texas A&M University at Galveston P.O. Box 1675 Galveston, Texas 77553 2 University of Texas Marine Science Institute University of Texas at Austin 750 Channel View Dr. Port Aransas, Texas 78383 3 Department of Zoology University of British Columbia 6270 University Boulevard Vancouver, BC, Canada V6T 1Z4. Declining populations of Atlantic sailfish (Istiophorus platypterus) (ICCAT, 2001) emphasize the need for a better understanding of their biology, especially processes affect- ing growth and mortality during early life because these mechanisms often regulate recruitment (Fuiman, 2002). To date, studies on sailfish biology during the early life interval are limited; however, recent research conducted in the Straits of Florida indicates that sailfish grow rapidly and experience high mortality during early life (Luthy et al., 2005; Rich- ardson et al., 2009a). Although our understanding of the early life ecol- ogy of sailfish has improved in recent years, work to date has been limited in geographic scope and basic life his- tory data on sailfish are limited or not available for other regions of the Atlantic that may represent essential spawning and nursery habitat of sail- fish. In particular, bycatch data from pelagic longline fisheries in the Gulf of Mexico indicate that catch rates for sailfish are twofold higher in the Gulf than in other areas of the North Atlantic during the presumed spawn- ing season (May-September; de Sylva and Breder, 1997; NMFS1). Because spawning stock biomass appears high in this region, a better understand- ing of the role of the Gulf of Mexico as spawning and early life habitat of Atlantic sailfish is essential. The Gulf of Mexico supports one of the most productive fisheries in the world (Chesney et al., 2000), and oceanographic features within the Gulf are influenced by inflow from the Mississippi River and large-scale oceanographic features, such as the Loop Current and associated warm and cold core eddies (Govoni and Grimes, 1992; Sturges and Leben, 2000). Physicochemical conditions and primary production vary mark- edly within and across these features (Grimes and Finucane, 1991; Biggs, 1992), and have been shown to af- fect the distribution and growth of pelagic larvae in the Gulf (Govoni et al., 1989; de Vries et al., 1990; Lang et al., 1994). Namely, higher densities 1 NMFS (National Marine Fisheries Ser- vice. 2008. NMFS Pelagic Longline Logbook data, (http://www.sefsc.noaa. gov/flslandingsdata.jsp, accessed Decem- ber 2008). Simms et al.: Distribution, growth, and mortality of larval Istiophorus plcitypterus in the northern Gulf of Mexico 479 and increased growth have been observed for larvae within frontal features created by riverine discharge and hydrodynamic convergence (Lang et ah, 1994; Hoff- meyer et ah, 2007). It is likely that oceanographic fea- tures also influence the early life ecology and population dynamics of sailfish in this region, and therefore an improved understanding of the effects of these features on distribution, growth, and survival will aid in deter- mining the value of the Gulf as spawning and nursery habitat for sailfish. The objectives of this research were threefold: 1) to characterize spatial and temporal patterns of abun- dance of sailfish larvae in the northern Gulf of Mexico; 2) to relate spatial variation in distribution and growth to oceanographic features to determine the causal fac- tors responsible for recruitment variability; and 3) to estimate demographic parameters within and across years to determine whether these traits varied tempo- rally. Estimates of growth (G) and mortality (Z) were combined ( G:Z ) to determine indices of stage-specific production potential and to assess the functional role of the Gulf as spawning and nursery habitat of this species. Materials and methods Field collections Five ichthyoplankton surveys were conducted in shelf and slope waters of the northern Gulf during spring and summer of 2005 (May, July, and September) and summer of 2006 (June and August). Surveys were con- ducted in an area bounded by 27° to 28°N latitude and 88 to 94°W longitude. This sampling area was selected because bycatch rates of adult sailfish by U.S. longliners were high in this region during the presumed spawning period of Atlantic sailfish (NMFS1). Istiophorid larvae were collected with paired neuston nets (2-m widthxl-m height frame) with two mesh sizes (500 pm and 1200 pm) to account for potential differences in capture success between mesh sizes. Nets were towed through the top meter of the water column at approximately 2.0 knots for 10 minutes. Paired tows were taken at 60 to 70 sampling stations spaced approximately 15 kilometers (km) apart during each survey. Sampling was conducted at 15-km intervals to allow coverage of a large area encompassing multiple oceanographic features. The September 2005 survey was shortened (39 stations sampled) because of inclement weather. At each station, sea surface temperature (°C), salin- ity (ppt), and dissolved oxygen (mg/L) were recorded by using a Sonde 6920 Environmental Monitoring Sys- tem (Yellow Springs Instruments Inc., Yellow Springs, OH). A probe malfunctioned during sampling in August 2006, preventing dissolved oxygen measurements. Sea surface height (cm) for each station was determined from archived satellite altimetry data provided by the Colorado Center for Astrodynamics Research (CCAR) Real-Time Altimetry Project (argo.colorado.edu; R. Leb- en, personal commun.2). Flowmeters (General Oceanics, model 2030R, Miami, FL) placed within each net were used to determine the surface area sampled by each net. A formula provided by the manufacturer was used to determine distance towed during each collection which was multiplied by net width to determine surface area sampled during each collection (m2): surface area sampled (m2) = distance sampled (m) x 2 m (net width). Fish larvae and associated biota were preserved on- board in 95% ethanol and istiophorids were sorted from each sample in the laboratory with the use of a Leica MZ stereomicroscope and stored in 70% ethanol. Istio- phorid larvae were photographed and standard length (SL) was measured to the nearest 0.1 mm. Istiophorids do not have a full complement of fin rays until reaching 20 mm SL (Richards, 1974), and 99.8% of specimens col- lected in the Gulf were less than 20 mm. Even though a small number of our largest specimens were over 20 mm and may be considered early juveniles in- 3), for the purposes of this study, all sailfish collected are referred to as larvae. Genetic identification A percentage of istiophorid larvae (22.2%) were iden- tified to the species level by following the protocol of Bangma (2006). The protocol was subsequently modi- fied and the remaining istiophorid larvae (77.8%) were identified according to the protocol of J. Magnussen and M. Shivji (personal commun.3). Briefly, a single eyeball was removed from each larva and DNA was extracted by using a QIAGEN DNeasy blood and tissue kit (QIAGEN #69506, Valencia, CA). A multiplex polymerase chain reaction (PCR) was performed by using an Eppendorf mastercycler gradient, QIAGEN Hot Star Taq DNA Polymerase (QIAGEN #203203), and PCR grade dNTP mix (QIAGEN #201901). Four primer pairs were used in each PCR reaction: a universal billfish primer set, and species-specific primers for sailfish, white marlin, ( Kajikia albida), and blue marlin ( Makaira nigricans). PCR reactions were examined by means of gel electro- phoresis with 1% agarose gels containing ethidium bro- mide and species identification was based on gel banding patterns. This multiplex assay was employed to identify sailfish larvae as follows. For samples consisting of less than 10 individuals, all specimens were assayed. For samples with 10-50 istiophorid larvae, 40—60% of larvae were processed. For samples with >50 istiophorid larvae, 25% of larvae were processed. If all larvae in a particu- lar subsample were identified as conspecific, remaining larvae from the sample were assigned to that species. If more than one istiophorid species was detected in a subsample, all remaining larvae from the sample were identified genetically. 2 Leben, Robert. 2008. Colorado Center for Astrodynamics Research, Univ. Colordao, Boulder, CO. 3 Shivji, Mahmood. 2007. Guy Harvey Research Institute, Nova Southeastern Univ., Ft. Lauderdale, FL. 480 Fishery Bulletin 108(4) Density and catch rate The total number of sailfish caught at each sampling station was divided by the surface area sampled to determine the density of larvae in number of individuals per 1000 m2. Mean density did not vary between the two mesh sizes (500 pm and 1200 pm) in any survey (paired t-test, all P>0.05), indicating no difference in capture success between net sizes. However, mean standard length was smaller in the 500-pm net gear (5.2 mm vs. 5.6 mm; Fa 3116)=21.3, PcO.Ol), indicating that a larger fraction of small larvae were retained by the finer mesh. Frequency of occurrence was calculated for each survey as the number of collection stations which produced at least one sailfish larva by using the equation Frequency of occurrence = ( number of stations with >1 larva / total number of (1) stations during survey) x 100. In order to compare catch rates with those from other ichthyoplankton studies, catch per unit of effort (CPUE) of sailfish larvae was estimated with the equation CPUE (larvae per hour) = number of larvae collected / ([ number of towsxtow length (min) x (2) 2 ( number of nets towed)] / 60). Oceanographic features Remotely sensed sea surface height (SSH) data were used to delineate oceanographic features into one of four categories (Leben et ah, 2002). Briefly, stations with a SSH greater than 17 cm were considered to be within the core of the Loop Current or an associated eddy system and classified as “anticyclones” (warm core eddies) (Leben et ah, 2002). Further, the core of adjacent “cyclones” (cold core eddies) were identified by a SSH of less than -10 cm. Frontal features associated with the Loop Current and anticyclones have been reported to extend up to 60 km from the 17-cm contour. Therefore, stations between 17 cm and -10 cm SSH and within 60 km of the 17 cm SSH contour were classified as being in a “front” (Tidwell, 2008). Remaining stations were classified as “open ocean.” Analysis of otolith microstructure Sagittal otoliths were extracted, cleaned of tissue in immersion oil, and preserved in mounting media (Flo- texx, Fisher Scientific #14-390-4, Pittsburgh, PA) for a subset of sailfish larvae. Mounted otoliths were pho- tographed under high magnification (400x) with an Olympus BX41 light microscope, and daily growth increments were enumerated with the use of Image- Pro Plus software (vers. 4.5, Media Cybernetics Inc., Bethesda, MD) by counting the first visible incre- ment beyond the hatch check as day one (Luthy et al. 2005). Inner increments of large otoliths are occasion- ally difficult to enumerate; therefore a regression of growth increment radius on age was used to predict the number of increments at various distances from the core (Rooker et al., 1999). The number of incre- ments within unclear regions was estimated with this regression and added to the increment count for the enumerated section to produce final age estimates. Less than 25% of age estimates were corrected with this method, and if the unclear region comprised more than 33% of the final age estimate, the otolith was not used for age and growth assessments. Two independent readings of daily increments were conducted by a single reader for each otolith. When two readings were within 10% variance of each other, one reading was randomly selected for analysis. If readings differed by >10%, a third reading was performed. If the third reading was within 10% variance of one of the former readings, one of the two similar readings was randomly selected for analysis. If each reading differed from the others by >10%, the otolith was not used for further analysis (n = 94, 7.1%). Growth, mortality and hatch dates Growth rates were determined by using otolith-derived age estimates (n = 1236) and standard length data. Because ages varied among surveys, growth rates were based on a limited age range (<17 days or the oldest larva in May 2005 collections) to minimize any effect of variable age ranges on growth analyses (Rilling and Houde, 1999). Daily instantaneous growth coefficients (g) were calculated from an exponential model: Lt = L0esf, (3) where Lt = length (mm SL) at time t; L0 = estimated length at hatching; g = instantaneous growth coefficient (d); and t - otolith-derived age (days after hatching). Ages of sailfish without an otolith-derived age were predicted by using age-length relationships (n=1118). A small number of sailfish larvae were damaged (n=72), and therefore length and age estimates could not be determined for these larvae. Daily mortality was estimated for each survey from regressions of the decline in log,, {abundances 1) on age. In order to minimize the influence of gear avoidance by larger larvae (Houde, 1987), mortality estimates were determined over a short time interval (10 days), and thus for a limited size range of larvae (<10 mm). Mor- tality was calculated for several time intervals rang- ing from 5 to 10 days and differences were negligible, indicating that the 10-day duration was appropriate for the target life stage. The age of peak abundance for each cohort was used as the initial point for catch-curve analysis and ranged from 9 to 11 days after hatching. Daily instantaneous mortality rates (Z) were calculated from the exponential model Simms et al.: Distribution, growth, and mortality of larval Istiophorus platypterus in the northern Gulf of Mexico 481 Table 1 Mean environmental conditions (± standard deviation [SD] ) by oceanographic feature for ichthyoplankton collections in the northern Gulf of Mexico in 2005 and 2006. Environmental parameters were recorded at the surface at each sampling station. Feature Temperature (°C) Salinity (ppt) Dissolved oxygen (mg/L) Anticyclone 29.1 (1.6) 36.1 (0.4) 6.8 (0.8) Front 28.9(1.4) 36.1 (0.4) 6.6 (0.5) Open ocean 29.3 (1.6) 35.5(1.2) 6.7 (0.8) Cyclone 28.2 (2.0) 35.6(0.6) 7.2 (1.3) Nt = N0e-Zt, (4) where Nt - abundance at time t\ N0 = estimated abundance at hatching; Z = instantaneous mortality coefficient (d); and t = otolith-derived age. Dry weight (mg) was calculated for all larvae with a measured length by using the length-weight relationship for sailfish larvae by Luthy et al. (2005): weight (mg) = 0.002(SL[mm] )3012. Weight-at-age data were fitted with exponential growth models to determine instantaneous weight-specific growth coefficients (G) for each survey by using the equation Wt = W0eGt, (5) where Wt = dry weight (mg) at time t ; W0 = estimated weight at hatching; G = instantaneous weight-specific growth coef- ficient (d); and t - otolith-derived age. Indices of stage-specific production potential were assessed for each cohort by examining the ratio of instantaneous weight-specific growth to daily mortal- ity ( G:Z ). This ratio incorporates growth and mortality and was used as an index of stage-specific production of larval cohorts (Rilling and Houde, 1999; Rooker et al., 1999; Wells et al., 2008). A cohort with a G.Z>1.0 was considered to be gaining biomass, which indicates that individuals had increased survival and production potential (Houde and Zastrow, 1993; Wells et al., 2008). Hatch dates for larvae were determined by sub- tracting age from date of collection. Otolith-derived ages were used when available, and remaining ages were predicted by applying cohort-specific age-length keys. Given that larger, older larvae in our collections hatched earlier and experienced greater cumulative mortality than larvae that hatched later, adjustments for mortality were made to more effectively represent the hatch dates of survivors in our collections by using the equation N0 = Nt / e~Zt, (Powell et al., 2004), (6) where N0 = estimated number of larvae at hatching; Nt = number of larvae at time t (Nt= 1 because N0 was calculated for each individual larva); Z = cohort-specific daily instantaneous mortality rate; and t = age of larva in days. Data analysis Spatial and temporal variation in environmental condi- tions and density of sailfish larvae was examined with a two-way analysis of variance (ANOVA) (factors: oceano- graphic feature and survey). Because of uneven repli- cates in 2005 and 2006, separate one-way ANOVAs were conducted to assess inter- and intra-annual variation in length and age of sailfish with year or survey as a fixed factor. In order to minimize heteroscedasticity, estimates of density were log^+1 transformed, whereas standard length and age data were loge transformed. In cases where variances were unequal, nonparametric analyses (Brown-Forsythe F-Test; Brown and Forsythe, 1974) were performed; however, results were consistent with para- metric tests (ANOVA) and thus only parametric analyses are presented. Post-hoc differences among levels of the main effect) s) were examined with Tukey’s honestly significant difference (HSD) test when variances were equal and with a Dunnett’s T3 test when variances were unequal (Zar, 1996). Analysis of covariance (ANCOVA) was used to test for spatial and temporal variations in growth and mortality (covariate: age) with models to determine if the slopes of the regression lines differed (slopes test). All data analyses were performed with SPSS, vers. 15.0 (SPSS Inc., Chicago, IL) with a=0.05. Results Environmental conditions Spatial and temporal variations in environmen- tal conditions were observed during Gulf collections. Temperatures were not significantly different among oceanographic features (ANOVA, F(3 270)= 2.2, P=0.09), albeit temperature was lowest within cyclones (28.2°C) compared with other oceanographic features (28.9- 29.3°C) (Table 1). Mean temperature was 28.8°C and 29.4°C in 2005 and 2006, respectively, and varied signifi- cantly among the five surveys (ANOVA, F{4 270)=354.0, 482 Fishery Bulletin 108(4) Table 2 Mean environmental conditions (± standard deviation [SD]) for ichthyoplankton surveys in the northern Gulf of Mexico in 2005 and 2006. Environmental parameters were recorded at the surface for each sampling station. Dissolved oxygen is not reported for the August 2006 survey because of a probe malfunction (NA). Year Survey Date Temperature (°C) Salinity (ppt) Dissolved oxygen (mg/L) 2005 May 17-22 26.4(0.9) 35.5 (0.7) 7.1 (0.7) July 23-28 30.4(0.7) 35.2(1.3) 6.8 (1.2) September 16-19 29.8(0.6) 36.5 (0.6) 6.3 (0.4) All surveys 28.8(2.0) 35.6 (1.1) 6.8 (1.0) 2006 June 15-20 28.7 (0.5) 35.8 (0.4) 6.6 (0.1) August 31 July- 5 Aug 30.1 (0.3) 36.2(0.4) NA All surveys 29.4(0.8) 36.0 (0.5) NA Table 3 Total catch, frequency of occurrence, and density (± standard deviation [SD] ) of sailfish ( Istiophorus platypterus ) larvae collected from the northern Gulf of Mexico in 2005 and 2006 arranged by survey. Number of stations during each survey (n) shown. Per- cent frequency of occurrence was based on collection stations during each survey that yielded 1 or more larvae. Survey n Sailfish catch Frequency of occurrence Density (± SD) (no. /1000m2) Maximum density Larvae/hour May 2005 60 212 26.7 0.6 (1.8) 10.5 10.6 July 2005 62 755 56.5 2.1 (7.5) 51.4 36.5 Sept 2005 39 134 46.2 0.6 (1.3) 4.5 10.3 June 2006 62 691 48.4 2.0 (4.7) 22.1 33.4 Aug 2006 65 634 47.0 1.7 (3.0) 10.2 25.4 Total 288 2426 45.0 1.5 (4.5) 24.4 PcO.Ol) (Table 2). Spatial variation in salinity was detected among oceanographic features (ANOVA, Pl3 270)=16.8, PcO.Ol), and lowest salinity observed within open ocean and cyclonic features (35.5 ppt and 35.6 ppt, respectively) and higher salinity in anticyclonic and fron- tal features (36.1 ppt and 36.1 ppt, respectively) (Table 1). Mean salinity was 35.6 ppt and 36.0 ppt in 2005 and 2006, respectively, and significant variation was observed among the five surveys (ANOVA, P(4 270) =11.1, PcO.Ol) (Table 2). Dissolved oxygen (DO) varied sig- nificantly among oceanographic features (ANOVA, P(3 270)=2.7, P=0.048) and surveys (ANOVA, P(4 270)=38.9, PcO.Ol) (Table 2). Lower DO levels were observed within anticyclones, fronts, and the open ocean (6.8 mg/L, 6.6 mg/L, and 6.7 mg/L, respectively) compared to those within cyclonic features (7.2 mg/L) (Table 1). Spatial and temporal patterns of abundance A significant interaction effect between oceanographic feature and survey on the density of sailfish was detected (ANOVA, F(10 271)=2.1, P=0.02), and highest densities were observed within frontal features in three of five surveys and in the open ocean during the remaining surveys (Fig. 1). No individual effect of oceanographic feature or survey on density was observed (feature P=0.06, survey P=0.11), although density was lowest within cyclones during all surveys. When data from all surveys were combined, significant spatial variation in the density of sailfish among oceanographic features was observed (ANOVA, P(3 285)=3.3, P=0.02); density within cyclones was lower than that within other fea- tures (P<0.05). Frequency of occurrence and relative abundances were highest in June (48.4%), July (56.5%), and August (47.0%) surveys and lowest during the May 2005 survey (26.7%) (Table 3). The overall density of sailfish was 1.5 larvae per 1000 m2; lowest densities were observed in May and September 2005 surveys (0.6 larvae per 1000 m2 and 0.6 larvae per 1000 m2, respectively) and higher densities during July 2005 (2.1 larvae per 1000 m2), June 2006 (2.0 larvae per 1000 m2), and August 2006 (1.7 larvae per 1000 m2) (Table 3). Catches peaked in July 2005 and June 2006 when maximum densities of 51.4 larvae per 1000 m2 and 22.1 larvae per 1000 m2 were observed, respectively (Table 3). The overall catch per unit of effort (CPUE) for sailfish in the northern Gulf was 24.4 larvae per hour, and highest CPUEs were reported for July 2005 (36.5), June 2006 (33.4), and August 2006 (25.4) (Table 3). Simms et al.: Distribution, growth, and mortality of larval Istiophorus platypterus in the northern Gulf of Mexico 483 Table 4 Exponential growth models arranged by survey and oceanographic feature for sailfish (Istiopho- rus platypterus) larvae collected from the northern Gulf of Mexico in 2005 and 2006. Number of larvae collected within each feature is given ( n ). Cyclones excluded from analysis because of low sample sizes within this type of feature. * indicates significant growth difference among features. Survey Feature n Growth model July 2005 Anticyclone 135 1.289e0128(ase) Front 45 1.458e° 122(agc) Open ocean 108 1.331e0124(age) June 2006 Anticyclone 21 1 7i3e° Front 108 1.415e° 120 August Anticyclone 131 1.682e0105(a^e,* 2006 Front 166 1.480e° u4(a«e) Open ocean 149 1.442e° 117(a^e)* Sailfish length and age distributions No significant difference in mean standard lengths of sailfish was observed between 2005 (5.1 mm, standard deviation [SD]=2.1) and 2006 (4.9 mm, SD=2.0) (ANOVA, F(1 2352)= 3.3, P-0.07) (Fig. 2). Sailfish were most abundant in the 3-6 mm size range, with 70.4% and 65.9% of the catch in this size range in 2005 and 2006, respectively. Intra- annual differences in mean lengths of sailfish were observed in 2005 (ANOVA, F(2 1034)=62.6, PcO.Ol) and 2006 (ANOVA, F(1 1315)=75.8, PcO.Ol); small- est mean lengths were observed in May 2005 (4.1 mm, SD-1.6) and June 2006 (4.5 mm, SD-1.6), and larger larvae were collected in July 2005 (5.4 mm, SD-2.2) and August 2006 (5.4 mm, SD-2.3). Mean ages were statistically similar between 2005 (10.1 days, SD-3.1 days) and 2006 (10.5 days, SD-3.2) (ANOVA, F(1 1234)=2.7, P-0.10) (Fig. 2). Sailfish larvae were most abundant in the 7-11 day range, with 60.7% and 53.9% of the catch in this age range in 2005 and 2006, respectively. Significant intra-annual variation in sailfish mean ages was observed in 2005 (ANOVA, P(2 521)=51.6, PcO.Ol) and 2006 (ANOVA, F(\ 7ioi=47.9, P<0.01), when youngest mean age was ob- served in May 2005 (8.0 days, SD-2.3) and June 2006 (9.4 days, SD-3.1), and older larvae were observed in July 2005 (11.0 days, SD-3.3) and August 2006 (11.0 days, SD-3.1). Spatial and temporal patterns of growth Spatial variation in growth was analyzed across oceano- graphic features (anticyclone, front, and open ocean) for three of the five surveys, and significant differences in 100 -i □ Anticyclone ■ Open ocean 0 Front □ Cyclone May 2005 July 2005 September June 2006 August surveys 2005 2006 100 w 2 03 to CO CD 2 ^ o a =3-4’ P-0.03) (Table 4). In contrast, growth did not differ significantly among the three features in July 2005 (anticyclones g=0.128, fronts g=0.122, and open ocean g= 0.124) and June 2006 (anticyclones g-0.101, fronts g-0.120, and open ocean g-0.112) (ANCOVA, 484 Fishery Bulletin 108(4) slopes, F{ 2 282)=®-^’ F=0.67 and ANCOVA, slopes, F(2 241)=2.5, P=0.09, respectively). Interannual differences in daily instantaneous growth rates of sailfish were detected between 2005 {g= 0.123) and 2006 (g=0.114) (ANCOVA, slopes, F(1 i205)=21.4, P<0.01) (Fig. 3). Further, significant intra-annual varia- tion in growth was observed in 2005 (ANCOVA, slopes, F[2 507) = 5.1, P=0.01); larvae collected in September (g- 0.113) displayed slower growth than larvae collected in July (g=0.127) (Fig. 4). In contrast, growth was sta- tistically similar between surveys in 2006 (ANCOVA, slopes, F(1 692)=0-7, P=0.39). Temporal variation in mortality Estimates of daily instantaneous mortality (Z) did not differ significantly between 2005 (Z=0.288) and 2006 (Z=0.394) (ANCOVA, slopes, P(1 46)=1.0, P=0.33) (Fig. 5, Table 5). Mortality rates ranged from 0.228 to 0.345 among survey periods in 2005, but rates were statisti- cally similar (ANCOVA, slopes, P(2 24)=2.1, P=0.15) (Fig. 6). Further, mortality rates were not significantly different between survey periods in 2006, ranging from 0.344 in June to 0.381 in August (ANCOVA, slopes, F(1 16)=0.5, P=0.48). Production potential (G.Z) Instantaneous weight-specific growth coefficients ( G ) were 0.371 in 2005 and 0.347 in 2006, and G was indexed to daily instantaneous mortality (Z) to assess stage- specific production potential, G.Z (Table 5). Annual estimates of G.Z were 1.29 in 2005 and 0.88 in 2006. Production potential was highest in May 2005 (1.30) and July 2005 (1.66). In contrast, G.Z ratios of the September 2005 (1.01), June 2006 (1.02), and August 2006 (0.91) surveys were lower. 400 300 200 ioo- 2005 10 120 100 - 80 - 60 - 40 25 400 300 200 ioo- 2006 120 - 80 60 - 40 - 20 0 5 10 15 20 25 Standard length (mm)/age (days) Figure 2 Standard length and age-frequency distributions of sailfish ( Istiopho - rus platypterus) larvae collected from the northern Gulf of Mexico in 2005 and 2006. Standard length distributions are shown in dark bars (n = 1037 and n = 1317 in 2005 and 2006, respectively). Age distributions are shown in light bars (n = 524 and n=712 in 2005 and 2006, respectively). Note: 72 larvae indicated in Table 2 were damaged and no length measurements were taken for these larvae. Hatch-date distribution Age-based estimates of hatch date indicated that sailfish in our collections were spawned from May to September (Fig. 7). Hatching of sailfish peaked in mid-July, and the majority of larvae hatched in July: 56.1% according to back-calculated estimates, 73.4% according to mortality adjusted estimates. The percent- age of total catch from July for mortality- adjusted hatch dates was 52.1% in 2005, and 77.1% in 2006. Because the majority of sailfish were <20 days of age and sampling was conducted bimonthly, hatch-date dis- tributions comprised multiple cohorts from separate spawning events throughout the spawning season. Discussion The relative abundance of sailfish larvae collected in the present study was compa- rable to or higher than values reported from other putative spawning grounds of sailfish. Llopiz and Cowen (2008), collected 7.3 sail- fish larvae per hour in neuston tows over a two-year period in waters of the Straits of Florida; their catch rate was lower than the catch rate of 27.4 sailfish larvae per hour reported by Post et al. (1997) in the same region. Further, Richardson et al. (2009b), during a single week, reported a catch rate of 298 sailfish larvae per hour in the Straits of Florida; this catch rate indicates that very high, but ephemeral, abundances occur in this region. Catch rates of istio- phorid billfishes (sailfish, white marlin, and blue marlin) reported in other areas, Simms et at: Distribution, growth, and mortality of larval Istiophorus platypterus in the northern Gulf of Mexico 485 including the Bahamas (Serafy et al., 2003) and the Dominican Republic (Prince et al., 2005), were less than 10 istiophorid larvae per hour, which is markedly lower than rates reported in our study for the Gulf of Mexico (24.4 sailfish larvae per hour). Sailfish frequency of occur- rence from the Gulf (45.0%) also corresponds closely to the 41.1% occurrence reported in the Straits of Florida and Bahamas by Luthy (2004), which included all Atlantic billfishes. Although CPUE was standardized for tow length, sampling gear, and number of tows, variations in towing methods as well as in the timing of sampling may have affected differences in catch rates of larvae. However, the relatively high catch rate of sailfish larvae reported in our study supports the premise that this region is an important spawn- ing and nursery ground of sailfish — a contention supported by high bycatch rates of adult sailfish during summer months (NMFS1). The highest density of larvae was reported within mesoscale frontal features and highest catch rates were observed from June through August. Lowest densities of sailfish larvae were observed in cold core features during all surveys; however, as with other pelagic species (Richards et al., 1993, Hoffmeyer et al., 2007), catches were higher within fronts and anticyclones associated with the western margin of the Loop Current. In fact, highest densities were observed within frontal features during three of five surveys and over the course of all surveys combined. Higher catches of marine fish larvae have been reported at frontal features created by riverine discharge or converging oceanic currents in both temper- ate and tropical oceans (Hoffmeyer et al., 2007; Richardson et al., 2009b). The accumulation of larvae near or within fronts may simply be due to hydrodynamic convergence which has been shown to aggregate marine larvae in the Gulf and other regions (Govoni and Grimes, 1992; Bakun, 2006). Alternatively, elevated primary and secondary production within frontal features often increases the availability of planktonic prey (Govoni et al., 1989; Grimes and Finucane, 1991) and therefore may effect higher survival for larvae within these oceanographic features (Grimes and Finucane, 1991; Biggs, 1992). In a recent study in the Straits of Florida, larval sailfish density was observed to peak at eddy frontal zones where there was a corresponding increase in density of common sailfish prey items (Llopiz and Cowen, 2008; Richardson et al., 2009b). This finding supports the premise that larvae are more abundant at fronts because of the increased availability of prey. As with spatial trends in sailfish density, temporal patterns in Gulf collections were comparable to those in the Straits of Florida; sailfish larvae were more abundant during spring and summer months (Post et al., 1997; Luthy, 2004). Elevated catches of sailfish in the summer correspond to peak spawning activity of sailfish in the North Atlantic (de Sylva and Breder, 1997; Richardson, 2007). Spatial variation in growth of sailfish was limited despite elevated densities of sailfish larvae within fronts. The general lack of growth variation in sailfish among oceanographic features is somewhat unexpected given that cyclonic and frontal features often display increased primary productivity relative to anticyclones (Grimes and Finucane, 1991; Biggs, 1992) and that growth variation during early life is influenced by pri- mary productivity and prey availability (de Vries et al., 1990; Wexler et al., 2007). The lack of variation in growth among oceanographic features may be at- tributed to the fact that larvae in the northern Gulf likely spend time in multiple oceanographic features during early life. Oceanographic currents in this re- gion have been observed to have speeds up to 0.8 me- ters per second (69.1 kilometers per day) (Govoni and Grimes, 1992), and higher velocity currents have been recorded within the Loop Current (1.0 meters per sec- 486 Fishery Bulletin 108(4) Figure 4 Size-at-age relationships of sailfish ( Istiophorus platypterus ) larvae collected from the northern Gulf of Mexico in 2005 and 2006 arranged by survey. Age in days was determined from otolith microstructure analysis. Exponential equations are given. Table 5 Instantaneous weight-specific growth ( G ) and mortality ( Z ) coefficients of sailfish ( Istiophorus platypterus ) larvae collected from the northern Gulf of Mexico in 2005 and 2006. Percentage per day was calculated from instantaneous growth and mortality coef- ficients. The G:Z index of stage-specific production potential is also shown. Year Survey G %/d ay Z %/day G:Z 2005 May 0.371 44.9 0.285 24.8 1.30 July 0.378 45.9 0.228 20.4 1.66 September 0.347 41.5 0.345 29.2 1.01 All surveys 0.371 44.9 0.288 25.0 1.29 2006 June 0.351 42.0 0.344 29.1 1.02 August 0.347 41.5 0.381 31.7 0.91 All surveys 0.347 41.5 0.394 32.6 0.88 Simms et al.: Distribution, growth, and mortality of larval Istiophorus platypterus in the northern Gulf of Mexico 487 2005 (n=524) LogeN, = 7.131 -0.288 (age) r2 = 0.94 o 0 10 15 20 2006 (n =71 2) Loge N, = 9.1 24 - 0.394(age) r2 = 0.95 10 Age (days) 15 Figure 5 Regression plots of loge (abundances 1) on age for ten-day cohorts of sailfish ( Istiophorus platypterus) larvae collected from the northern Gulf of Mexico in 2005 and 2006. Regression equations and plots are arranged by year. ond; Vukovich and Maul, 1985), which indicate that planktonic larvae may encounter multiple oceanographic features during their planktonic larval duration. Further, pelagic larvae from broad oceanic areas have been reported to ac- cumulate within frontal zones (Richards et al., 1993; Bakun, 2006), making it difficult to de- termine where individuals spend the majority of their lives and therefore, in which feature(s) most of their early growth occurs. Estimated growth of sailfish in the Gulf (g=0.113 to 0.127) varied temporally and rates were comparable to or slightly slower than those reported for sailfish in the Straits of Florida (g=0.130 to 0.146; Luthy et al., 2005; Richardson, 2007; Richardson et al., 2009a) and blue marlin from the Bahamas (g= 0.098 to 0.125; Serafy et al., 2003; Sponaugle et al., 2005) and the Straits of Florida (g=0.089 to 0.114; Sponaugle et al., 2005; Richardson, 2007). Observed differences in growth among studies are minor and similarities are not un- expected because the timing of collections and environmental conditions between the regions were comparable. Sampling in the Straits of Florida was conducted between April and September in waters ranging from 26.1°C to 30.6°C (Luthy et al., 2005; Richardson, 2007). This range of temperatures is similar to tem- peratures present in the Gulf during our May to September sampling period (26.4-30.4 °C). Moreover, observed ranges in salinity were nearly the same between the Straits of Flor- ida (34.0-36.7 ppt) and Gulf (35.2-36.5 ppt). Temporal variations in growth of marine lar- vae have been shown to be correlated with temperature (Rilling and Houde, 1999), and the most rapid growth of sailfish larvae was observed during July 2005 when the warmest mean temperature was reported. Nevertheless, the second fastest growth rate for sailfish was observed during May 2005 when the lowest mean temperature was reported. Thus, other factors known to affect growth, such as density of con- specifics or prey availability (Jenkins et al., 1991; Lang et al., 1994; Wexler et al., 2007), may be responsible for observed variations in growth of sailfish larvae. Although intra-annual differences in mortality were limited, losses were substantial throughout the early life interval examined (Z = 0.23 to 0.38). Daily instan- taneous mortality rates reported in our study are 10-45% lower than mortality rates for sailfish and blue marlin larvae from the Straits of Florida and Exuma Sound, Bahamas (Richardson et al., 2009a). However, the losses reported in our study are compa- rable to those of other pelagic larvae, such as bluefin tuna ( Thunnus thynnus ) (Z=0.20; Rooker et al., 2007), yellowfin tuna ( Thunnus albacares) (Z= 0.33; Lang et al., 1994), and members of the suborder Scombroidei, which includes tunas, billfishes, and the barracuda Sphyraena barracuda (Z = 0.34; Houde and Zastrow, 1993). Predation has been observed to be a major cause of mortality during the early life interval of pe- lagic species (Leggett and Deblois, 1994; Houde, 2002) and it may be responsible for high losses of sailfish larvae. Recent studies indicate that istiophorids are preyed upon by conspecifics and congeners (Llopiz and Cowen, 2008; Tidwell, 2008) and, if so, cannibalism or predation pressure by other istiophorids may represent an important source of mortality for sailfish during early life. Observed G:Z ratios were greater than 1.0 during all but the August 2006 survey, indicating conditions were likely favorable for production during the life stage examined. Houde and Zastrow (1993) reported a mean G:Z ratio of 0.89 for marine fish larvae (pooled across several taxa), which is lower than the range reported in our study for sailfish (0.91-1.66). However, indices reported for individual species or taxonomic groups ranged from 0.26 to 2.42 for larvae abundant in upwelling and shelf zones, indicating wide-ranging stage-specific production potential for pelagic fishes. Collection of larvae peaked in July and it is possible that increased G:Z coincided with increased temper- ature or prey availability, both of which have been 488 Fishery Bulletin 108(4) August 2006 (n=459) Loge A/, = 8.645 - 0.381 (age) r2 = 0.94 Ns o o n 10 12 14 16 18 20 Age (days) Figure 6 Regression plots of logp {abundance + 1 ) on age for ten-day cohorts of sailfish (Istiopho- rus platypterus) larvae collected from the northern Gulf of Mexico in 2005 and 2006. Regression equations and plots arranged by survey. shown to increase stage-specific production potential (Rilling and Houde, 1999). Hatch dates of sailfish larvae collected in this study indicate that spawning is protracted in the northern Gulf and peak activity occurs during July. To date, sailfish spawning has not been documented in the northern Gulf, but sailfish are known to have pro- tracted spawning in other regions of the Atlantic and Pacific oceans (de Sylva and Breder, 1997; Chiang et al., 2006; Richardson, 2007). Although the timing of sampling likely influences hatch-date distributions of sailfish in the Gulf, resulting in a multimodal distri- bution, spawning has been shown to occur from May to September in the western North Atlantic Ocean (de Sylva and Breder, 1997) and from April to Sep- tember in the eastern Pacific (Chiang et ah, 2006), corresponding to the spawning range observed in this study. Additionally, data from studies indicate that peak spawning occurs in mid to late summer because increased frequencies of mature ovaries have been ob- served in sailfish landed in July and August (de Sylva and Breder, 1997; Chiang et ah, 2006; Richardson et al., 2009a). Questions remain regarding the specific factors regu- lating observed variations in distribution, growth, and mortality of sailfish, as well as the extent of spawn- ing in the northern Gulf. Regardless, high densities and broad distributions of larvae combined with rapid growth and high production potential indicate that sailfish larvae spawned or hatched in the northern Gulf potentially contribute to Atlantic sailfish populations. This study provides strong evidence that the northern Gulf serves as an important spawning and nursery habitat for Atlantic sailfish. Simms et al.: Distribution, growth, and mortality of larval Istiophorus platypterus in the northern Gulf of Mexico 489 -10.0 “0 CD O