[ pe rates Ree marae parent hts flor ay net 1 Aree PT i) Hg ay bee unica ie le Hh way H fata J elite Pia hata syn ; i i a —————— CeCe ele. Vg ATOLL RESEARCH BULLETIN NOS. 459-465 RESEARCH BULLETIN ed by TIONAL MUSEUM OF NATURAL HISTORY THSONIAN INSTITUTION SHINGTON, D.C. U.S.A. susT 1999 \) - ATOLL RESEARCH BULLETIN NOS. 459-465 NO. NO. NO. NO. NO. NO. . 459. 460. 461. 462. 463. 464. 465. SPECIES RICHNESS OF RECENT SCLERACTINIA BY STEPHEN D. CAIRNS ATOLLS AS SETTLEMENT LANDSCAPES: UJAE, MARSHALL ISLANDS BY MARSHALL I. WEISLER REPORT ON FISH COLLECTIONS FROM THE PITCAIRN ISLANDS BY JOHN E. RANDALL FISH NAMES IN LANGUAGES OF TONGA AND FIJI BY R. CHRISTOPHER MORGAN THE NON-NATIVE VASCULAR PLANTS OF HENDERSON ISLAND, SOUTH CENTRAL PACIFIC OCEAN BY STEVE WALDREN, MARSHALL I. WEISLER, JON G. HATHER AND DYLAN MORROW REVISED VEGETATION CLASSIFICATION OF TURNEFFE ATOLL, BELIZE BY MALCOLM R. MURRAY, SIMON A ZISMAN AND CHRISTOPHER D. MINTY A MICROBIALITE/ALGAL RIDGE FRINGING REEF COMPLEX, HIGHBORNE CAY, BAHAMAS BY R. PAMELA REID, IAN G. MACINTYRE AND ROBERT S. STENECK ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. AUGUST 1999 ACKNOWLEDGMENT The Atoll Research Bulletin is issued by the Smithsonian Institution to provide an outlet for information on the biota of tropical islands and reefs and on the environment that supports the biota. The Bulletin is supported by the National Museum of Natural History and is produced by the Smithsonian Press. This issue is partly financed and distributed with funds from Atoll Research Bulletin readers and authors. The Bulletin was founded in 1951 and the first 117 numbers were issued by the Pacific Science Board, National Academy of Sciences, with financial support from the Office of Naval Research. Its pages were devoted largely to reports resulting from the Pacific Science Board's Coral Atoll Program. All statements made in papers published in the Atoll Research Bulletin are the sole responsibility of the authors and do not necessarily represent the views of the Smithsonian nor of the editors of the Bulletin. Articles submitted for publication in the Atoll Research Bulletin should be original papers in a format similar to that found in recent issues of the Bulletin. First drafts of manuscripts should be typewritten double spaced and can be sent to any of the editors. After the manuscript has been reviewed and accepted, the author will be provided with a page format with which to prepare a single-spaced camera-ready copy of the manuscript. COORDINATING EDITOR Ian G. Macintyre National Museum of Natural History MRC-125 ASSISTANTS Smithsonian Institution Kasandra D. Brockington Washington, D.C. 20560 William T. Boykins, Jr. Thomas C. Niemann EDITORIAL BOARD Stephen D. Cairns (MRC-163) National Museum of Natural History Brian F. Kensley (MRC-163) (Insert appropriate MRC code) Mark M. Littler (MRC-166) Smithsonian Institution Wayne N. Mathis (MRC-169) Washington, D.C. 20560 Jeffrey T. Williams (MRC-159) Joshua I. Tracey, Jr. (MRC-137) Warren L. Wagner (MRC-166) Roger B. Clapp National Museum of Natural History National Biological Survey, MRC-111 Smithsonian Institution Washington, D.C. 20560 David R. Stoddart Department of Geography 501 Earth Sciences Building University of California Berkeley, CA 94720 Bernard M. Salvat Ecole Pratique des Hautes Etudes Labo. Biologie Marine et Malacologie Université de Perpignan 66025 Perpignan Cedex, France PUBLICATIONS MANAGER A. Alan Burchell Smithsonian Institution Press ATOLL RESEARCH BULLETIN NO. 459 SPECIES RICHNESS OF RECENT SCLERACTINIA BY STEPHEN D. CAIRNS ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. AUGUST 1999 ees (noe Jems eee ae <<>> MT Aer ‘e there mete Mk - » — SPECIES RICHNESS OF RECENT SCLERACTINIA BY STEPHEN D. CAIRNS' ABSTRACT Most previous estimates of the number of valid, described Recent species (known species richness) of Scleractinia have been unsupported guesses ranging from 1000-2500 species. The actual number, based on a list of all senior synonyms, is approximately 1314, classified in 24 families and 220 genera (average number of species per genus = 5.97), 79 of the genera being monotypic. The numbers of zooxanthellate and azooxanthellate species and genera are about the same: e.g., zooxanthellates contributing to 48.2% of the genera and 49.5% of the species. Over the last three decades an average of 16.1 new species of Scleractinia have been described each year. Although the yearly rate of new descriptions is very uneven, the decadal trend appears to indicate a gradual decrease in the number of newly described zooxanthellate species and genera, balanced by an increase in the number of newly described azooxanthellate species and genera. An estimate of total species richness was made based on the perceived ratio of described to undescribed species of Scleractinia ascertained from the analysis of comprehensive faunistic analyses and taxonomic revisions. This method estimates a minimum of 1479 species. A second, less reliable method, which is based on the rates of species descriptions over time, suggests a range of 1460-2628 species. EPIGRAPH “There are about 2500 living species of corals and over 5000 extinct ones; hence these animals reached their height in past ages and are now on the decline.” (Hyman, 1940: 620) KNOWN SPECIES RICHNESS Historical Estimates Estimates of the number of valid, described, living (modern) species of Scleractinia range from a low of 1000 (Kaestner, 1964) to a high of 2500 species (Hyman, 1940)(see Table 1). Most of these estimates are educated guesses, not accompanied or based on a listing of actual species names that would allow for hypothesis testing and constructive criticism. The first publication purporting to list all scleractinian species was that of the World Conservation 'Department of Invertebrate Zoology, National Museum of Natural History, Smithsonian Institution, Washington, D.C., 20560-0163 Manuscript received 6 November 1998; revised 7 April 1999 2. Table 1._Estimates of Known and Total Species Richness of Scleractinia Year Author Zooxanthellate AZOox- Total I-Pac Atl. Total anthellate Estimates of Known Species Richness: 1925: 825 Kiikenthal - - - - 2500 1940:620 Hyman . - - - 2500 1943: 77-90 Vaughan & Wells >500 48 548 ~453 >1001 1956: 360 Wells >500 - - - - 1967:115 Kaestner . - - - 1000 1967:79 Wilmoth - - - - 2500 1981:120 Rosen 500 68 568 - - 1982:611 Cairns & Stanley - . 940 560 1500 1982:701 Dunn - - - - 2500 1985:37 Naumoy, et al. - - 55079 = 2500 1985: 18 Kiihlmann - - - - 2500 1986: 179 Cairns, et al. - - - - 2500 1986:1 Veron ~500 - - - - 1987:642,668 Chevalier 700 70 TION e<850 <1620 1988: 67 Schuhmacher 500 684 584anh- - 1989: 35 Zibrowius 500 60 560 ~560 ~1120 1990: 206 Brusca & Brusca - - - - 2500 1991: 476 Jackson - - 750 =- - 1993: 60-136 WCMC 547 68 615 425 1040 1995: 160 Veron - - 833i - 1997: 2 Cairns . - - 617 - 1999 (herein) Cairns, Hoeksema & van der Land 585 70 656 669 *1314 *allows for 11 facultative species Estimates of Total Species Richness: Based on partial inventory (see text) - - >696 >781 >1479 Based on rate of description (see text) 1460-2628 Table 2._Numbers of valid species (and genera), monotypic genera, and average number of species per genus of the Recent Scleractinia, arranged by family from highest number of species to lowest. Ave. Family Zooxan- Azooxan- *Facul- Total Mono- Species thellate thellate tative typic Per Genera Genus Caryophylliidae 25(10) 274(43) 3(2) 296(51) 17 5.8 Acroporidae 199(4) 0 0 199(4) 0 49.8 Dendrophylliidae 15(3) 135(17) 2(1) =148(19) 4 7.8 Faviidae 103(24) 0 0 103(24) 9 4.3 Flabellidae 0 98(10) 0 98(10) 2 9.8 Poritidae 74(4) 0 0 74(4) l 18.5 Turbinoliidae 0 51(22) 0 51(22) iT 2.3 Mussidae 46(13) 0 0 46(13) 6 3.5 Agariciidae 45(7) 0 0 45(7) 3 6.4 Fungiidae 44(11) 0 0 44(11) 3 4.0 Rhizangiidae 1(1) 33(4) 1d) 33(4) 0 8.3 Pocilloporidae 22(4) 10(1) 2(1) 30(4) 1 TES Siderastreidae 27(6) 0 0 27(6) 3 4.5 Oculinidae 14(5) 15(6) 3(1) =26(10) o/ 2.6 Fungiacyathidae 0 20(1) 0 20(1) 0 20.0 Pectinidae 19(5) 0 0 19(5) 0 3.8 Micrabaciidae 0 13(4) 0 13(4) 0 3.3 Merulinidae 12(5) 0 0 12(5) 3 2.4 Anthemiphylliidae 0 7(1) 0 7(1) 0 7.0 Guyniidae 0 7(7) 0 7(7) 7 1.0 Gardineriidae 0 5(1) 0 5(1) 0 5.0 Meandrinidae 5(4) 0 0 5(4) 3 123 Astrocoeniidae 4(2) 0 0 4(2) 1 2.0 Trachyphylliidae 1(1) 0 0 1(1) 1 1.0 Incertae Sedis 0 1(-) 0 1(-) - - TOTALS: 656(109) 669(117) 11(6) 1314(220) 79 5:97 *Facultative: Eleven species may occur in the zooxanthellate and azooxanthellate forms: three species of Heterocyathus, two species of Heteropsammia, two species of Madracis, Astrangia poculata, and three species of Oculina. These species are counted as both zooxanthellates and azooxanthellates, but only once in the total column. Cladocora also contains species, some of which are exclusively zooxanthellate, others exclusively azooxanthellate. 4 Monitoring Centre (WCMC, 1993), compiled by E. Wood for the purpose of listing all scleractinian species regulated by CITES (Convention on International Trade in Endangered Species of Wild Fauna and Flora). Although a worthy first attempt, this listing of 1040 species is considered to be flawed (uncritical) in that it used outdated taxonomy, occasionally included fossil species and genera, included some junior synonyms, and included some duplication of names. It also employed a confusing, three-tiered system of categorizing species, i.e., nominal, valid, and “well-established”, and was far from complete regarding the azooxanthellate species. The only other listing known to include all Recent scleractinian species was an unpublished draft (1995) of 1259 species submitted at the Sixth International Conference on Coelenterate Biology (ICCB VI) as part of a larger series included in the Unesco-IOC Register of Marine Organisms (ed. J. van der Land, 1995). It is that list, which is herein corrected and updated, that forms the basis for the 1314 species listed in the Appendix. The Current Number The current (end of 1998) number of 1314 valid, Recent scleractinian species (see Appendix), is summarized by family in Table 2. From this compilation one can see that there are 24 families of Recent Scleractinia, containing 220 genera, 79 (36%) of which are monotypic. Twelve families are exclusively zooxanthellate, 7 families are exclusively azooxanthellate, and 5 families contain genera having both ecological classes. In fact, eleven species and six genera are facultative, existing in both the zooxanthellate or azooxanthellate forms, depending on the environment (listed as a footnote to Table 2). The numbers of zooxanthellate and azooxanthellate species is virtually the same, 656 (49.5%) vs 669 (50.5%), respectively. Likewise, the number of genera is almost the same, with only a slight majority favoring the azooxanthellates at both taxonomic levels. There is an average of 5.97 and a range of 1-49.8 species per genus. Since Cairns (1997) calculated a similar species average of 5.37 for the exclusively azooxanthellate species, this ratio is approximately the same for both zooxanthellates and azooxanthellates. It should be stated at this point that the species concept used in this paper is the morphospecies, or operational morphotaxonomic unit (sensu Veron, 1995), first formalized for scleractinian corals by Vaughan (1907: 4) as: “...a group of individuals connected among themselves by intergrading characters and separated by distinct lacunae from all other individuals or groups of individuals.” Molecular, physiological, behavioral, and ecological evidence of species distinction (see Lang, 1984) were undoubtedly used to help construct the list of coral species, but these kind of data are not currently available for many species and not at all for fossil species, which makes the morphospecies most appropriate when comparing faunas within the fossil record. The “species problem” in corals is amply discussed by Veron (1995), including ramifications of reticulate evolution and the philosophy of conceptual vs operational species definitions. But, for the purposes of this paper, the traditional morphospecies is employed. RATES OF NEW SPECIES DESCRIPTIONS Over the last 30.5 years (1968 to mid-1998) the rate of description of new species of Scleractinia has been very uneven (Figure 1), reflecting the aperiodic publication of major faunistic revisions. (No judgment of the validity of these newly described species is made herein.) In this time interval, 490 species were described, or an average of 16.1 species per year. The uneveness of the yearly description totals is reflected in a range of 1-56 and a rather high standard deviation of 12.3. However, when viewed on a decadal scale (Table 3), some trends are apparent. Whereas the number of new species descriptions has seemed to reach a plateau over the last 20 years, the zooxanthellate:azooxanthellate components have altered dramatically. There appears to be a decline in both the number of newly described species and genera of zooxanthellate corals, which is replaced by an increase in the number of both genera and species of azooxanthellate corals. In fact, no new genera of zooxanthellate were described in the last decade, whereas 20 new azooxanthellate genera were described in the same time period. This change in rates of description might suggest that the more accessible, shallow-water zooxanthellate species are becoming fairly well known worldwide, especially at the generic level, whereas the primarily deep-water azooxanthellate fauna is less well known and thus might contribute more to the increase in described scleractinian species richness in the future. 60 : : = ee : | a Figure 1. Number of new species of Scleractinia | —-@ -Zooxanthellete | described/year (1968 - 1997) ---A--- Azooxanthellete |_| f | | =o Total | | Average Number of Species Described ESTIMATES OF TOTAL SPECIES RICHNESS Armed with these statistics on the known fauna, it is tempting to predict the total species richness of Scleractinia. One method of estimating global species richness in a taxon is the partial inventory method, which relies on the perceived ratio of described to undescribed species ascertained by a specialist in that group and/or by analysis of the literature. For instance, over the last 20 years the average percentage of previously undescribed azooxanthellate species in 14 faunistic studies from 12 regions (Cairns, 1979, 1982, 1984, 1989, 1991, 1994, 1995, 1998, 1999, in press a; Cairns & Parker, 1992; Cairns & Keller, 1993; Cairns & Zibrowius, 1997; Zibrowius, 1980) was 14.3% (range = 5.0-24.0%), or conversely 85.7% previously described. If this average described ratio is assumed to apply to the entire currently known azooxanthellate fauna (669 + 0.857), one might expect there to be 781 azooxanthellate species worldwide. If similar logic is applied to zooxanthellate corals, a smaller ratio of 6.1% (range 0-18.2%) undescribed, or conversely, 93.9% previously described species results. This undescribed:described ratio is based on the following 15 publications covering six regions and two taxonomic revisions: eastern Australia (Veron et al., 1976-1984); western Australia (Veron, 1985; Veron & Marsh, 1988); Japan (Veron, 1990, 1992); Viet Nam (Latypov, 1990, 1992); Red Sea (Scheer & Pillai, 1983); Caribbean (Zlatarski, 1982); family Fungiidae (Hoeksema, 1989); and genus Leptoseris (Dinesen, 1980). This average described ratio applied to the known zooxanthellate fauna (656 + 0.939), results in the prediction of 698 species. Together, the zooxanthellate and azooxanthellate estimates total 1479 (Table 1). To reiterate, the assumptions implicit in this estimation are: 1) 1314 currently known valid species, composed of 656 zooxanthellates and 669 azooxanthellates, and 2) a minimal undescribed component of 14.3% for azooxanthellates and 6.1% for zooxanthellates. Table 3.—Decadal trends in rates of description of species (and genera) of Recent zooxanthellate and azooxanthellate Scleractinia. *Average for second decade corrected because Zoological Record volume 123 covered 1.5 years, making total period analyzed 30.5 years. Zoological Years Zooxan- Azooxan- Total Ave. Number Growth Overall Record of thellate thellate Species/Y ear Rate (%) Rate Volume Coverage 105-114 1968-77 100(6) 29(0) 129(6) 12.9 1.35 2.97 115-124 1978-87/88 114(9) 69(8) 183(17) iii 1.53 Bi ay? 125-134 1988-97/98 65(0) —-113(20) 178(20) 17.8 1.35 Bi TOTAL: 1968-97/98 279(15) 211(28) 490(43) 16.1 1.23 2.94 A second method of estimating diversity, developed by Hammond (1992), is based on an analysis of time series of species description rates. First, one calculates the current growth rate per annum of the taxon in question, i.e., the number of species described per year divided by the total number of valid species. Using the average number of scleractinian species described per year over the last 30 years (16.1) and the current total number of scleractinian species, this equation is 16.1 + 1314, or 1.23%, implying that over the last 30 years the number of scleractinian species increased by about 1.23%/year, although due to synonymy this percentage is certainly lower. Decadal rates are also given in Table 3. Secondly, Hammond calculates the ratio of the current rate = overall rate, the overall rate being the average yearly rate of species descriptions since 1758. Again, using the average number of scleractinian species described per year over the last 30 years (16.1) and the overall rate of 1314 species + 240 years (=1998-1758), yields the equation: 16.1 + (1314+240), or 2.94, implying that over the last 30 years corals have been described at 2.94 times the post-Linnaean “average rate.” Decadal rates are also listed in Table 3. Hammond then compares these two ratios (the growth rate and current rate/overall rate) with the ratios derived for other animal groups (which for scleractinian corals is coincidentally the same as that for fish), and rather subjectively designates a value for “the proportion of species described to date.” According to Hammond these two ratios are consistent with taxa having a “high” proportion of previously described species, i.e., 50-90%. Applying this percentage to 1314 species results in an estimation of 1460-2628 species (see Table 2). Assumptions implicit in this estimation are: 1) 1314 currently valid species, 2) all newly described species are valid, and 3) acceptance of implications of species growth rates and overall rates as intuited by Hammond (1992). DISCUSSION Methods for estimating global species richness of various taxa are highly controversial, often conflicting, and usually difficult to apply. Useful reviews on this topic include: May (1990), Hammond (1992, 1994), Stork (1993, 1997), and Colwell & Coddington (1994). Some of these methods rely on the principle of taxon ratios, wherein a reference site is chosen for which one element of the fauna is thought to be fairly well known (or at least well sampled), providing an estimate of the described:undescribed species ratio for that taxon for that site. This ratio is then applied to the currently known species richness of a larger area that includes the reference site (hierarchical taxon ratio) or a separate geographic area (non-hierarchical taxon ratio) to obtain estimates of species richness. The first method used in this paper, the “partial inventory method,” falls into this category and is patterned, in large part, on a study by Hodkinson & Casson (1991), who attempted to determine global insect biodiversity using a hierarchical taxon ratio. After extensive sampling of Hemiptera in northern Sulawesi, Hodkinson & Casson determined that 62.5% of the collected species were undescribed. Then, making family-by-family comparisons of Sulawesi to world species, they showed that the same proportion of new species is likely to be found worldwide. Using these ratios and the currently known species richness of Hemiptera, they were able to provide a reasonable estimate of the worldwide Hemiptera species richness. Critics of this method (Stork, 1993, 1997; Hammond, 1994) point out that it is virtually impossible to claim that all species, whether insect or corals, are known from any reference site, regardless of the intensity of collection. This is a valid criticism, and for that reason the estimates that result from such studies should be considered as minimum estimates. A second criticism of this method is the assumption that the 8 described:undescribed ratio of one well sampled area is representative of the rest of the world. To ameliorate this criticism, I have chosen an average described:undescribed ratio from 18 regions and two taxonomic revisions, and furthermore established two different ratios, one for zooxanthellates and the other for azooxanthellates. Hammond’s method of using trends in description rate to predict global species diversity has been criticized by Erwin (1991) and Hammond (1992) himself, and is not a frequently used estimator for species richness. Rates of description depend on many factors, including one’s species concept, the number of taxonomists working on a group at any period of time, and the technology used to investigate species. There also appears to be a bias to describe species of large body size and for which material is available, more commonly from temperate localities. Finally, Hammond’s classification of the “proportion of species described to date” is extremely subjective (intuitive) and essentially undefined (unscientific). Also, the influence of new technology on known species richness is unpredictable. For instance, molecular analysis (allozymes) has suggested an increase in the number of Montastraea sibling species (Knowlton et al., 1992), whereas similar techniques have suggested a reduction in the number of recognized Platygyra species (Miller & Benzie, 1997); however, molecular data “have generally been found to support traditional morphological interpretations of species boundaries” (Wallace & Willis, 1994: 248; see also Willis, 1990). Synonymy of species is also a common result of more thorough morphological examination of larger suites of specimens from more diverse areas. Thus, the tendencies to increase the number of known species (e.g., discovery of sibling species) are often offset by the synonymy of species based on morphological and/or molecular methods. The overall effect is impossible to predict. Although Hammond’s method is highly subjective and rarely used, it is one of the few methods available to predict total scleractinian species richness and does suggest, in my opinion, a reasonable range. On a purely intuitive basis, I would estimate the total number of scleractinian species to be about 2100, implying that we have described about 63% of the known fauna and that about 790 species remain to be described in this order. CONCLUSIONS What drives some people to want to know how many species exist on this planet or, more specifically, how many species occur in a particular taxon? The traditional answers are usually threefold (May, 1990; Stork, 1993). A knowledge of species richness: 1) helps establish a necessary first step to understand how biological systems work and provides a baseline that would allow for their conservation (ecology and conservation argument), 2) allows for the potential use of a greater variety of species for pharmaceutical products (utilitarian argument), and 3) satisfies the simple, unadulterated curiosity to know (quixotic argument). In addition to these traditional arguments, I would suggest that knowing the actual number of Recent scleractinian species is a valuable reference point for comparisons to late Tertiary faunas (one might call the Recent Benchmark or Paleontological Baseline argument). Hyman’s (1940) conclusion that corals are “on the decline” because there are now only 2500 living species and 5000 extinct ones, is incorrect and misleading in many ways. First, there are far fewer than 2500 valid living species; it is absurd to compare taxa from the Recent to the entire Phanerozoic; comparing Mesozoic-Tertiary Scleractinia to Paleozoic Rugosa and Tabulata is illogical; and finally the total number of reef scleractinian corals appears to vacillate in time (Budd, in press) and is not a simple trend. And yet this ill-founded statement seems to have influenced two generations of textbook writers and even coral biologists (Table 1). According to Veron (1995: figs. 25, 36-38) and Scrutton (1997: fig. 2), Scleractinia stand at an all time maximum of generic diversity, but can the same be said at the species level? In very thorough studies of the Caribbean Neogene zooxanthellate Scleractinia, Budd, Stemann & Johnson (1994, Table 5) and Budd, Johnson & Stemann (1996) found 67-100% more species throughout the Late Miocene to Pliocene at 2 MY intervals than in the Recent, whereas Cairns (in press a, b) found considerably more Recent azooxanthellate species (131 species) than in the comparable Caribbean Neogene (49 species). Thus, whether the Caribbean zooxanthellates are on the decline and the azooxanthellates are on the increase, or whether the latter assumption 1s due to the artefact of “the pull of the Recent,” it is essential to have an accurate baseline figure of Recent species richness to even begin these or similar speculations. ACKNOWLEDGEMENTS I would like to thank the following people who read earlier drafts of this manuscript and provided valuable advice and suggestions: Nancy Budd, Vladimir Kosmynin, Bert Hoeksema, and Timothy Werner. I also thank Peter Glynn for access to his advance list of reef corals from the eastern Pacific. REFERENCES Brusca, RC, Brusca, GJ (1990) Invertebrates. Sinauer Associates, Inc., Sunderland, 922 pp. Budd, AF (in press) Diversity and extinction in the Cenozoic history of Caribbean reefs. In: Hughs, TP , Sale, P (eds) Ecology of coral reefs: biological and evolutionary approaches. Yale University Press. Budd, AF, Johnson, KG, Stemann, TA (1996) Plio-Pleistocene turnover and extinctions in the Caribbean reef-coral fauna. In: Jackson, JBC, Budd, AF, Coates, AG (eds) Evolution and Environment in Tropical America. University of Chicago Press, Chicago, pp. 168-204. Budd, AF, Stemann, TA, Johnson, KG (1994) Stratigraphic distributions of genera and species of Neogene to Recent Caribbean reef corals. J Paleont 68(5): 951-977. Cairns, SD (1979) The deep-water Scleractinia of the Caribbean Sea and adjacent waters. Stud Fauna Cura¢gao 57(180): 341 pp. Cairns, SD (1982) Antarctic and Subantarctic Scleractinia. Antarct Res Ser 34(1): 74 pp. Cairns, SD (1984) New records of ahermatypic corals (Scleractinia) from the Hawaiian and Line Islands. Occas Pap Bernice Pauahi Bishop Mus 25(10): 1-30. Cairns, SD (1989) A revision of the ahermatypic Scleractinia of the Philippine Islands and adjacent waters, Part 1: Fungiacyathidae, Micrabaciidae, Turbinoliinae, Guyniidae, and Flabellidae. Smithson Contrib Zool 486: 136 pp. 10 Cairns, SD (1991) A revision of the ahermatypic Scleractinia of the Galapagos and Cocos Islands. Smithson Contrib Zool 504: 32 pp. Cairns, SD (1994) Scleractinia of the temperate North Pacific. Smithson Contrib Zool 557: 150 Pp. Cairns, SD (1995) The marine fauna of New Zealand: Scleractinia (Cnidaria: Anthozoa). N Z Oceanogr Inst Mem 103: 139 pp. Cairns, SD (1997) A generic revision and phylogenetic analysis of the Turbinoliidae (Cnidaria: Scleractinia). Smithson Contrib Zool 591: 55 pp. Cairns, SD (1998) Azooxanthellate Scleractinia (Cnidaria: Anthozoa) of Western Australia. Rec West Australian Mus 18: 361-417. Cairns, SD (1999) Cnidaria Anthozoa; deep-water azooxanthellate Scleractinia from Vanuatu, and Wallis and Futuna Islands. Mém Mus natn Hist nat 180: 31-167. Cairns, SD (in press, a) A revision of the shallow-water azooxanthellate Scleractinia of the western Atlantic. Stud. nat. Hist. Caribb region. Cairns, SD (in press, b) Stratigraphic distribution of Neogene Caribbean azooxanthellate corals (Scleractinia and Stylasteridae). Bull Am Paleont Cairns, SD, den Hartog, JC, Arneson, C (1986) Class Anthozoa. In: Sterrer, W. (ed) Marine Fauna and Flora of Bermuda. John Wiley & Sons, New York, pp. 159-194. Cairns, SD, Keller, NB (1993) New taxa and distributional records of azooxanthellate Scleractinia from the tropical south-west Indian Ocean, with comments on their zoogeography and ecology. Ann S Afr Mus 103(5): 213-292. Cairns, SD, Parker, SA (1992) Review of the Recent Scleractinia of South Australia, Victoria, and Tasmania. Rec S Austr Mus, Monogr Ser 3: 82 pp. Cairns, SD, Stanley, GD (1982) Ahermatypic coral banks: living and fossil counterparts. Proc 4th Int Coral Reef Symp 1: 611-618. Cairns, SD, Zibrowius, H (1997) Cnidaria Anthozoa: azooxanthellate Scleractinia from the Philippine and Indonesian regions. Mém Mus natn Hist nat 172: 27-243. Chevalier, J-P (1987) Ordre de Scléractiniaires: Systématique. In: Grassé, P-P (ed) Traité de Zoologie 3(3) Masson, Paris, pp. 679-753. Colwell, RK, Coddington, JA (1994) Estimating terrestrial biodiversity through extrapolation. Phil Trans R Soc London B 345: 101-118. Dinesen, ZD (1980) A revision of the coral genus Leptoseris (Scleractinia: Fungiina: Agariciidae). Mem Qd Mus 20(1): 181-235. Dunn, DF (1982) Cnidaria. In: Parker, SP (ed.) Synopsis and Classification of Living Organisms. McGraw-Hill Book Co., New York, pp 669-706. Erwin, TL (1991) How many species are there?: revisited. Conserv Biol 5: 330-333. Hammond, PM (1992). Species inventory. In: Broombridge, B (ed) Global Diversity: Status of the Earth's Living Resources. Chapman & Hall, London, pp. 17-39. Hammond, PM (1994) Practical approaches to the estimation of the extent of biodiversity in speciose groups. Phil Trans R Soc London B 345: 119-136. Hodkinson, ID, Casson, D (1991) A lesser predilection for bugs: Hemiptera (Insecta) diversity in tropical forests. Biol J Linn Soc 43: 101-109. Hoeksema, BW (1989) Taxonomy, phylogeny, and biogeography of mushroom corals 11 (Scleractinia: Fungiidae). Zool Verh (Leiden) 254: 295 pp. Hyman, LH (1940) The Invertebrates: Protozoa through Ctenophora. McGraw-Hill Book Company, New York, 726 pp. Jackson, JBC (1991) Adaptation and diversity of reef corals. Bioscience 41(7): 475-482. Kaestner, A. (1967) Invertebrate Zoology, Volume 1. Interscience Publishers, New York, 597 pp. Knowlton, N, Weil, E, Weigt, LA, Guzman, HM (1992) Sibling species in Montastraea annularis, coral bleaching, and the coral climate record. Science 255: 330-333. Kukenthal, W (1925) Madreporaria In: Krumbach, T (ed.) Handbuch der Zoologie 1, Walter de Gruyter & Co., Berlin, p. 825.. Kiihlmann, DHH (1985) Living Coral Reefs of the World. Arco Publishing, Inc, New York, 185 Pp. Land, van der, J (ed)(1995, unpublished draft) Unesco-IOC Register of Marine Organisms: Coelenterata or Cnidaria, Class Hexacorallia: Order Scleractinia (Stony Corals), National Museum of Natural History, Leiden, 21 pp. Lang, JC (1984) Whatever works: the variable importance of skeletal and of non-skeletal characters in scleractinian taxonomy. Paleont Amer 54: 18-44. Latypov, YA (1990) Scleractinia corals from Viet Nam, Part 1: Thamnasteriidae, Astrocoeniidae, Pocilloporidae, Dendrophylliidae. Hayka, Moscow, 81 pp. [In Russian] Latypov, YA (1992) Scleractinian corals from Viet Nam, Part 2: Acroporidae. Hayka, Moscow, 131 pp. [In Russian] May, RM (1990) How many species? Phil Trans R Soc (B)330: 293-304. Miller, KJ, Benzie, JAH (1997) No clear cut genetic distinction between morphological species within the coral genus Platygyra. Bull Mar Sci 61(3): 907-917. Naumov, DB, Propp, MB, Ribakov, CH, 1985. World of Corals. Gidrometeoizdat, Leningrad, 359 pp. [in Russian] Rosen, DR (1981) The tropical high diversity enigma - the coral's-eye view. In: Forey, PL (ed) The Evolving Biosphere, Cambridge University Press, Cambridge, pp. 103-129. Scheer, G, Pillai, CSG (1983) Report on the stony corals from the Red Sea. Zoologica 133: 198 pp. Schuhmacher, H (1988) Korallenriffe. BLV Verlagsgesellschaft, Miinchen, 275 pp. Scrutton, CT (1997) The Palaeozoic corals, I: origin and relationships. Proc Yorkshire Geol Soc 51(3): 177-208. Stork, NE (1993) How many species are there? Biodiv Conserv 2: 215-232. Stork, NE (1997) Measuring global biodiversity and its decline. In: Reaka-Kudla, ML, Wilson, DE, Wilson, EO (eds) Biodiversity II. John Henry Press, Washington, DC, pp. 41-68. Vaughan, TW (1907) Recent Madreporaria of the Hawaiian Islands and Laysan. Bull US Nat Mus 59: 427 pp. Vaughan, TW, Wells, JW (1943) Revision of the suborders families, and genera of the Scleractinia. Geol Soc Amer Spec pap 44: 363 pp. Veron, JEN (1985) New Scleractinia from Australian coral reefs. Rec West Aust Mus 12(1): 147-183. Veron, JEN (1986) Corals of Australia and the Indo-Pacific. Angus & Robertson Publishers, North Ryde, 644 pp. 12 Veron, JEN (1990) New Scleractinia from Japan and other Indo-West Pacific countries. Galaxea 9:95-173. Veron, JEN (1992) Hermatypic corals of Japan. Aust Inst Mar. Sci Monogr 9: 234 pp. Veron, JEN (1995) Corals in Space and Time. UNSW Press, Sydney, 321 pp. Veron, JEN, Marsh, LM (1988) Hermatypic corals of Western Australia. Rec West Aust Mus, Supplement 29: 136 pp. Veron, JEN, et al. (1976-1984) Scleractinia of Eastern Australia: Parts 1-5. Aust Inst Mar Sci Monogr Ser 1, 3-6: 1383 pp. Wallace, CC, Willis, BL (1994) Systematics of the coral genus Acropora: implications of new biological findings for species concepts. Annu Rev Ecol Syst 25: 237-262. Wells, JW (1956) Scleractinia. In: Moore, RC (ed) Treatise on Invertebrate Paleontology, Part F, Coelenterata. Geological Society of America, Lawrence, Kansas, pp. F328-F444. Willis, BL (1990) Species concepts in extant scleractinian corals: considerations based on reproductive biology and genotypic population structures. Syst Bot 15(1): 136-149. Wilmoth, JH (1967) Biology of Invertebrata. Prentice Hall, Inc., Englewood Cliffs, 465 pp. World Conservation Monitoring Centre (1993) Checklist of fish and invertebrates listed in the CITES appendices. Joint Nature Conservation Committee, Peterborough, 171 pp. Zibrowius, H (1980) Les Scléractiniaires de la Méditerranée et de |'Atlantique nord-oriental. Mém Inst Océanogr (Monaco) 11: 284 pp. Zibrowius, H (1989) Mise au point sur les Scléractiniaires comme indicateurs de profondeur. Géol Médit 15(1): 27-47. Zlatarski, VN (1982, French ed.) Description systématique. In: Zlatarski, VN, Estalella, NM, Les Scléractiniares de Cuba. Acad Bulgare Sci, Sofia, pp. 25-343. Appendix: List of Extant Stony Corals BY STEPHEN D. CAIRNS!, BERT W. HOEKSEMA? and JACOB VAN DER LAND? PREFACE Within the five cnidarian orders that contain species having calcified skeletons (i.e., the "stony corals"), all valid, extant species are listed below in alphabetical order by family, genus, and species. Their general distribution is indicated to the right of each species by numbers: 1, western Atlantic; 2, eastern Atlantic; 3, Indian Ocean; 4, western and central Pacific; 5, eastern Pacific; and 6, Subantarctic and Antarctic regions. A question mark in a column indicates a questionable occurrence in this region. For the scleractinians, the azooxanthellate species are marked with an asterisk, the zooxanthellate are unmarked, and those 11 species that occur as both forms are marked with a cross (+). This is believed to be the first complete and critical listing of all 1574 species of extant stony corals, consisting of 1314 scleractinians and 260 calcified hydrozoans. It is meant to complement a similar Internet version of the same list to be released as part of the UNESCO-IOC Register of Marine Organisms (ed., J. van der Land, National Museum of Natural History, Leiden), the first draft of which was compiled in 1995. Although every effort was made to make the list as complete and accurate as possible through 1998, we acknowledge that there are certainly errors of omissions and interpretation. We consider this as a first effort to establish a data base of all valid, extant species, and welcome any comments and corrections to the list. In general, the first author was responsible for the accuracy of the species included in the azooxanthellate Scleractinia, western Atlantic Scleractinia, and Stylasteridae, whereas the second author was responsible for the Indo-West Pacific zooxanthellate Scleractinia and calcified hydrozoans (Milleporidae). We acknowledge that there exist other calcified octocorallian cnidarians that are not listed herein, pertaining to the families: Tubiporidae, Helioporidae, Lithotelestidae, Coralliidae, and Isididae, as well as calcified hydrozoans of the genus Pseudsolandaria. We believe the value of such a list to be manifold. It serves as a documentation of the species richness of larger taxa; it provides authorship and date of publication of all species; it provides a starting point for identification of corals from various geographic regions; it serves as 'Department of Invertebrate Zoology, National Museum of Natural History, Smithsonian Institution, Washington, D. C. 20560 *National Museum of Natural History, Postbus 9517, 2300 RA Leiden, The Netherlands 14 a point of reference before a new species is described; and it may help to avoid homonymy in future described species, although one should be aware that fossil species are not listed in this account. Phylum CNIDARIA Class ANTHOZOA Subclass HEXACORALLIA Order SCLERACTINIA Distribution Acroporidae Acropora abrolhosensis Veron, 1985 3 Acropora aculeus (Dana, 1846) Acropora acuminata (Verrill, 1864) Acropora akajimensis Veron, 1990 Acropora anthocercis (Brook, 1893) Acropora arabensis Hodgson & Carpenter, 1996 Acropora aspera (Dana, 1846) Acropora austera (Dana, 1846) Acropora awi Wallace & Wolstenholme, 1998 Acropora azurea Veron & Wallace, 1984 Acropora batunai Wallace, 1997 Acropora branchi Riegl, 1995 3 Acropora brueggemanni (Brook, 1893) 3 Acropora bushyensis Veron & Wallace, 1984 3 Acropora cardenae Wells, 1986 Acropora carduus (Dana, 1846) 3 Acropora caroliniana Nemenzo, 1976 3 Acropora cerealis (Dana, 1846) 3 Acropora cervicornis (Lamarck, 1816) ] Acropora chesterfieldensis Veron & Wallace, 1984 Acropora clathrata (Brook, 1891) 3 Acropora copiosa Nemenzo, 1967 Acropora crateriformis (Gardiner, 1898) Acropora cuneata (Dana, 1846) Acropora cytherea (Dana, 1846) Acropora danai (Milne Edwards & Haime, 1860) Acropora dendrum (Bassett-Smith, 1890) Acropora derawaensis Wallace, 1997 Acropora desalwii Wallace, 1994 Acropora digitifera (Dana, 1846) 3 Acropora divaricata (Dana, 1846) 3 Acropora donei Veron & Wallace, 1984 3 Acropora echinata (Dana, 1846) 3 Acropora elegans (M. Edwards & Haime, 1860) Acropora elseyi (Brook, 1892) 3 Acropora eurystoma (Klunzinger, 1879) 3 Acropora exquisita Nemenzo, 1971 3 WWW WwW WwW oS ee Fh HHoHL WWW WwW fHHH HH hh HHH HH HHH HHHAHSA P= Acropora florida (Dana, 1846) Acropora formosa (Dana, 1846) Acropora gemmifera (Brook, 1892) Acropora glauca (Brook, 1893) Acropora grandis (Brook, 1892) Acropora granulosa (M. Edwards & Haime, 1860) Acropora halmaherae Wallace & Wolstenholme, 1998 Acropora hemprichii (Ehrenberg, 1834) Acropora hoeksemai Wallace, 1997 Acropora horrida (Dana, 1846) Acropora humilis (Dana, 1846) Acropora hyacinthus (Dana, 1846) Acropora indiana Wallace, 1994 Acropora indonesia Wallace, 1997 Acropora insignis Nemenzo, 1967 Acropora intermedia (Brook, 1891) Acropora jacquelinae Wallace, 1994 Acropora kirstyae Veron & Wallace, 1984 Acropora kosurini Wallace, 1994 Acropora latistella (Brook, 1892) Acropora listeri (Brook, 1893) Acropora loisetteae Wallace, 1994 Acropora lokani Wallace, 1994 Acropora longicyathus (Milne Edwards & Haime, 1860) Acropora loripes (Brook, 1892) Acropora lovelli Veron & Wallace, 1984 Acropora lutkeni Crossland, 1952 Acropora magnifica Nemenzo, 1971 Acropora microclados (Ehrenberg, 1834) Acropora microphthalma (Verrill, 1869) Acropora millepora (Ehrenberg, 1834) Acropora mirabilis Quelch, 1886 Acropora monticulosa (Brueggemann, 1879) Acropora mossambica Riegl, 1995 Acropora multiacuta Nemenzo, 1967 Acropora nana (Studer, 1878) Acropora nasuta (Dana, 1846) Acropora natalensis Riegl, 1995 Acropora nobilis (Dana, 1846) Acropora ocellata (Klunzinger, 1879) Acropora palifera (Lamarck, 1816) Acropora palmata (Lamarck, 1816) Acropora palmerae Wells, 1954 Acropora paniculata Verrill, 1902 Acropora parilis Quelch, 1886 Acropora pharaonis (M. Edwards & Haime, 1860) Acropora plumosa Wallace & Wolstenholme, 1998 Acropora pocilloporina Wallace, 1994 Acropora polystoma (Brook, 1891) Acropora prolifera (Lamarck, 1816) Acropora pruinosa (Brook, 1893) Ww WW WW WwW WwW WWW WW WwW Ww WWWW WW WW WwW WW WWW WW WWW WW WwW Ww AAP HPAP HH HPA HAHAHAHAHAHA RS FHA HHH H HALA AS _ 15 16 Acropora pulchra (Brook, 1891) Acropora rambleri Bassett-Smith, 1890 Acropora robusta (Dana, 1846) Acropora rosaria (Dana, 1846) Acropora rudis (Rehberg, 1892) Acropora russelli Wallace, 1994 Acropora samoensis (Brook, 1891) Acropora sarmentosa (Brook, 1892) Acropora schmitti Wells, 1950 Acropora secale (Studer, 1878) Acropora sekiseiensis Veron, 1990 Acropora selago (Studer, 1878) Acropora simplex Wallace & Wolstenholme, 1998 Acropora solitaryensis Veron & Wallace, 1984 Acropora sordiensis Riegl, 1995 Acropora spicifera (Dana, 1846) Acropora squarrosa (Ehrenberg, 1834) Acropora stoddarti Pillai & Scheer, 1976 Acropora striata (Verrill, 1866) Acropora subglabra (Brook, 1891) Acropora subulata (Dana, 1846) Acropora suharsonoi Wallace, 1994 Acropora sukarnoi Wallace, 1997 Acropora tanegashimensis Veron, 1990 Acropora tenella (Brook, 1892) Acropora tenuis (Dana, 1846) Acropora teres (Verrill, 1866) Acropora togianensis Wallace, 1997 Acropora torihalimeda Wallace, 1994 Acropora tortuosa (Dana, 1846) Acropora tumida Verrill, 1866 Acropora turaki Wallace, 1994 Acropora valenciennesi (M. Edwards & Haime, 1860) Acropora valida (Dana, 1846) Acropora vaughani Wells, 1954 Acropora verweyi Veron & Wallace, 1984 Acropora wallacea Veron, 1990 Acropora willisae Veron & Wallace, 1984 Acropora yongei Veron & Wallace, 1984 Anacropora forbesi Ridley, 1884 Anacropora matthai Pillai, 1973 Anacropora puertogalerae Nemenzo, 1964 Anacropora reticulata Veron & Wallace, 1984 Anacropora spinosa Rehberg, 1892 Astreopora cucullata Lamberts, 1980 Astreopora explanata Veron, 1985 Astreopora gracilis Bernard, 1896 Astreopora incrustans Bernard, 1896 Astreopora lambertsi Moll & Best, 1984 Astreopora listeri Bernard, 1896 Astreopora macrostoma Veron & Wallace, 1984 Wo WWW WwW WwW WWW WwW Ww WWW Ww WW WWW WW WH WH WH WW Ww WW WwW WwW WwW WwW S$HHHH HHH HHP HHHHL P= ee LPHRHKH HA HHA HAHAHAHA HH HAH HHA HHA HHH Astreopora moretonensis Veron & Wallace, 1984 Astreopora myriophthalma (Lamarck, 1816) Astreopora ocellata Bernard, 1896 Astreopora suggesta Wells, 1954 Montipora aequituberculata Bernard, 1897 Montipora altasepta Nemenzo, 1964 Montipora angulata (Lamarck, 1816) Montipora australiensis Bernard, 1897 Montipora cactus Bernard, 1897 Montipora calcarea Bernard, 1897 Montipora caliculata (Dana, 1846) Montipora capitata Dana, 1846 Montipora capricornis Veron, 1985 Montipora cebuensis Nemenzo, 1976 Montipora circumvallata (Ehrenberg, 1834) Montipora confusa Nemenzo, 1967 Montipora corbettensis Veron & Wallace, 1984 Montipora crassituberculata Bernard, 1897 Montipora danae (M. Edwards & Haime, 1851) Ww PHA HHP HHH HAHA HSA Montipora digitata (Dana, 1846) Montipora edwardsi Bernard, 1879 Montipora efflorescens Bernard, 1897 Montipora effusa Dana, 1846 Montipora florida Nemenzo, 1967 Montipora floweri Wells, 1954 Montipora foliosa (Pallas, 1766) Montipora foveolata (Dana, 1846) Montipora friabilis Bernard, 1897 Montipora gaimardi Bernard, 1897 Montipora granulosa Bernard, 1897 Montipora grisea Bernard, 1897 Montipora hirsuta Nemenzo, 1967 Montipora hispida (Dana, 1846) Montipora hoffmeisteri Wells, 1954 Montipora incrassata (Dana, 1846) Montipora informis Bernard, 1897 Montipora lobulata Bernard, 1897 Montipora mactanensis Nemenzo, 1979 Montipora malampaya Nemenzo, 1967 Montipora millepora Crossland, 1952 Montipora mollis Bernard, 1897 Montipora monasteriata (Forskal, 1775) Montipora nodosa (Dana, 1846) Montipora orientalis Nemenzo, 1967 Montipora peltiformis Bernard, 1897 Montipora samarensis Nemenzo, 1967 Montipora setosa Nemenzo, 1976 Montipora solanderi Bernard, 1879 Montipora spongiosa (Ehrenberg, 1834) Montipora spongodes Bernard, 1897 Montipora spumosa (Lamarck, 1816) WWW WWW WWW WWW WH Ww WWW WwW WwW Wo WWW WwW WWW Ww S&H HHH pH HH HAHAHAHAHAHA ppHHHK HAHAHA a 18 Montipora stellata Bernard, 1897 Montipora stilosa (Ehrenberg, 1834) Montipora striata Bernard, 1897 Montipora tuberculosa (Lamarck, 1816) Montipora turgescens Bernard, 1897 Montipora turtlensis Veron & Wallace, 1984 Montipora undata Bernard, 1897 Montipora venosa (Ehrenberg, 1834) Montipora verrucosa (Lamarck, 1816) Agariciidae Agaricia agaricites (Linnaeus, 1758) Agaricia fragilis Dana, 1846 Agaricia grahamae Wells, 1973 Agaricia humilis Verrill, 1902 Agaricia lamarcki M. Edwards & Haime, 1851 Agaricia tenuifolia Dana, 1846 Agaricia undata (Ellis & Solander, 1786) Coeloseris mayeri Vaughan, 1918 Gardineroseris planulata (Dana, 1846) Helioseris cucullata (Ellis & Solander, 1786) Leptoseris amitoriensis Veron, 1990 Leptoseris cailleti (Duchassaing & Michelotti, 1864) Leptoseris explanata Yabe & Sugiyama, 1941 Leptoseris foliosa Dinesen, 1980 Leptoseris gardineri Van der Horst, 1921 Leptoseris hawaiiensis Vaughan, 1907 Leptoseris incrustans (Quelch, 1886) Leptoseris mycetoseroides Wells, 1954 Leptoseris papyracea (Dana, 1846) Leptoseris scabra Vaughan, 1907 Leptoseris solida (Quelch, 1886) Leptoseris tenuis Van der Horst, 1921 Leptoseris tubulifera Vaughan, 1907 Leptoseris yabei (Pillai & Scheer, 1976) Pachyseris foliosa Veron, 1990 Pachyseris gemmae Nemenzo, 1955 Pachyseris rugosa (Lamarck, 1801) Pachyseris speciosa (Dana, 1846) Pavona bipartita Nemenzo, 1980 Pavona cactus (Forskal, 1775) Pavona clavus (Dana, 1846) Pavona danai (M. Edwards & Haime, 1816) Pavona decussata (Dana, 1846) Pavona diffluens Lamarck, 1816 Pavona divaricata Lamarck, 1816 Pavona duerdeni Vaughan, 1907 Pavona explanulata (Lamarck, 1816) Pavona frondifera Lamarck, 1816 Pavona gigantea Verrill, 1869 Pavona lata Dana, 1846 ] WWW WW WWW Ww WwW Wo WWW WWW WWW WW WW WwW WWW WWW WW WwW HHH HHL pss AHHH HH HHA HH HPP HHH HH pp HoH HHH Pavona maldivensis (Gardiner, 1905) Pavona minuta Wells, 1954 Pavona varians Verrill, 1864 Pavona venosa (Ehrenberg, 1834) Pavona xarifae Scheer & Pillai, 1974 Anthemiphylliidae *Anthemiphyllia dentata (Alcock, 1902) *Anthemiphyllia frustum Cairns, 1994 * Anthemiphyllia macrolobata Cairns, 1998 *Anthemiphyllia multidentata Cairns, 1998 *Anthemiphyllia pacifica Vaughan, 1907 *Anthemiphyllia patera patera De Pourtalés, 1878 *A. patera costata Cairns, 1999 *Anthemiphyllia spinifera Cairns, 1999 Astrocoeniidae Stephanocoenia intersepta (Lamarck, 1816) Stylocoeniella armata Ehrenberg, 1834 Stylocoeniella cocosensis Veron, 1990 Stylocoeniella guentheri Bassett-Smith, 1890 Caryophylliidae *Anomocora carinata Cairns, 1991 *Anomocora fecunda (De Pourtalés, 1871) *Asterosmilia gigas (van der Horst, 1931) * Asterosmilia marchadi (Chevalier, 1966) *Asterosmilia prolifera (De Pourtalés, 1871) *Aulocyathus atlanticus Zibrowius, 1980 *Aulocyathus juvenescens Marenzeller, 1904 *Aulocyathus matricidus (Kent, 1871) *Aulocyathus recidivus (Dennant, 1906) *Bathycyathus chilensis M. Edwards & Haime, 1848 *Bourneotrochus stellulatus (Cairns, 1984) *Caryophyllia abrupta Cairns, 1999 *Caryophyllia abyssorum Duncan, 1873 *Caryophyllia alaskensis Vaughan, 1941 *Caryophyllia alberti Zibrowius, 1980 *Caryophyllia ambrosia ambrosia Alcock, 1898 *C. ambrosia caribbeana Cairns, 1979 *Caryophyllia antarctica Marenzeller, 1904 *Caryophyllia antillarum De Pourtalés, 1874 *Caryophyllia arnoldi Vaughan, 1900 *Caryophyllia atlantica (Duncan, 1873) *Caryophyllia balanacea Zibrowius & Gili, 1990 *Caryophyllia barbadensis Cairns, 1979 *Caryophyllia berteriana Duchassaing, 1850 *Caryophyllia calveri Duncan, 1873 *Caryophyllia capensis Gardiner, 1904 *Caryophyllia cincticulatus (Alcock, 1898) N i) tO WWW Ww Ww Ww ARP HHA PHP HAA BSS .o) 19 20 *Caryophyllia cornulum Cairns & Zibrowius, 1997 *Caryophyllia corrugata Cairns, 1979 *Caryophyllia crosnieri Cairns & Zibrowius, 1997 *Caryophyllia cyathus (Ellis & Solander, 1786) *Caryophyllia decamera Cairns, 1998 *Caryophyllia dentata Moseley, 1876 *Caryophyllia diomedeae Marenzeller, 1904 *Caryophyllia eltaninae Cairns, 1982 *Caryophyllia ephyala Alcock, 1891 *Caryophyllia foresti Zibrowius, 1980 *Caryophyllia grandis Gardiner & Waugh, 1938 *Caryophyllia grayi (M. Edwards & Haime, 1848) *Caryophyllia hawaiiensis Vaughan, 1907 *Caryophyllia horologium Cairns, 1977 *Caryophyllia inornata (Duncan, 1878) *Caryophyllia japonica Marenzeller, 1888 *Caryophyllia jogashimaensis Eguchi, 1968 *Caryophyllia karubarica Cairns & Zibrowius, 1997 *Caryophyllia lamellifera Moseley, 1881 *Caryophyllia mabahithi Gardiner & Waugh, 1938 *Caryophyllia marmorea Cairns, 1984 *Caryophyllia octonaria Cairns & Zibrowius, 1997 *Caryophyllia octopali Vaughan, 1907 *Caryophyllia paradoxus Alcock, 1898 *Caryophyllia paucipalata Moseley, 1881 *Caryophyllia pauciseptata Yabe & Eguchi, 1932 *Caryophyllia perculta Cairns, 1991 *Caryophyllia planilamellata Dennant, 1906 *Caryophyllia polygona De Pourtalés, 1878 *Caryophyllia profunda Moseley, 1881 *Caryophyllia quadragenaria Alcock, 1902 *Caryophyllia quangdongensis Zou, 1984 *Caryophyllia ralphae Cairns, 1995 *Caryophyllia rugosa Moseley, 1881 *Caryophyllia sarsiae Zibrowius 1974 *Caryophyllia scillaemorpha Alcock, 1894 *Caryophyllia scobinosa Alcock, 1902 *Caryophyllia secta Cairns & Zibrowius, 1997 *Caryophyllia seguenzae Duncan, 1873 *Caryophyllia smithii Stokes & Broderip, 1828 *Caryophyllia solida Cairns, 1991 *Caryophyllia spinicarens (Moseley, 1881) *Caryophyllia spinigera Saville Kent, 1871 *Caryophyllia squiresi Cairns, 1982 *Caryophyllia stellula Cairns, 1998 *Caryophyllia transversalis Moseley, 1881 *Caryophyllia unicristata Cairns & Zibrowius, 1997 *Caryophyllia valdiviae Zibrowius & Gili, 1990 *Caryophyllia zanzibarensis Zou, 1984 *Caryophyllia zopyros Cairns, 1979 Catalaphyllia jardinei (Saville-Kent, 1893) wo = HHH HH HH *Ceratotrochus franciscana Durham & Barnard, 1952 *Ceratotrochus magnaghii Cecchini, 1914 Cladocora arbuscula Lesueur, 1881 Cladocora caespitosa (Linnaeus, 1758) *Cladocora debilis M. Edwards & Haime, 1849 *Cladocora pacifica Cairns, 1991 *Coenocyathus anthophyllites M. Edwards & Haime, 1848 *Coenocyathus bowersi Vaughan, 1906 *Coenocyathus brooki Cairns, 1995 *Coenocyathus cylindricus M. Edwards & Haime, 1848 *Coenocyathus goreaui Wells, 1972 *Coenocyathus parvulus (Cairns, 1979) *Coenosmilia arbuscula De Pourtalés, 1874 *Coenosmilia inordinata Cairns, 1984 *Colangia immersa De Pourtalés, 1871 *Colangia moseleyi (Faustino, 1927) *Concentrotheca laevigata (De Pourtalés, 1871) *Concentrotheca vaughani Cairns, 1991 *Confluphyllia juncta Cairns & Zibrowius, 1997 *Conotrochus asymmetros Cairns, 1999 *Conotrochus brunneus (Moseley, 1881) *Conotrochus funicolumna (Alcock, 1902) *Crispatotrochus cornu (Moseley, 1881) *Crispatotrochus curvatus Cairns, 1995 *Crispatotrochus foxi (Durham & Barnard, 1952) *Crispatotrochus galapagensis Cairns, 1991 *Crispatotrochus inornatus Tenison-Woods, 1878 *Crispatotrochus irregularis (Cairns, 1982) *Crispatotrochus niinoi (Yabe & Eguchi, 1942) *Crispatotrochus rubescens (Moseley, 1881) *Crispatotrochus rugosus Cairns, 1995 *Crispatotrochus squiresi (Cairns, 1979) *Crispatotrochus woodsi (Wells, 1964) *Dactylotrochus cervicornis (Moseley, 1881) *Dasmosmilia lymani (De Pourtalés, 1871) *Dasmosmilia valida (Marenzeller, 1907) *Dasmosmilia variegata (De Pourtalés, 1871) *Deltocyathus agassizi De Pourtalés, 1867 *Deltocyathus andamanicus Alcock, 1898 *Deltocyathus calcar De Pourtalés, 1874 *Deltocyathus cameratus Cairns, 1999 *Deltocyathus corrugatus Cairns, 1999 *Deltocyathus crassiseptum Cairns, 1999 *Deltocyathus eccentricus Cairns, 1979 *Deltocyathus halianthus (Lindstrém, 1877) *Deltocyathus heteroclitus Wells, 1984 *Deltocyathus italicus (Michelotti, 1838) *Deltocyathus magnificus Moseley, 1876 *Deltocyathus moseleyi Cairns, 1979 *Deltocyathus murrayi Gardiner & Waugh, 1938 *Deltocyathus ornatus Gardiner, 1899 iw) NO BSH HHH 22 *Deltocyathus parvulus Keller, 1982 *Deltocyathus philippinensis Cairns & Zibrowius, 1997 *Deltocyathus pourtalesi Cairns, 1979 1 *Deltocyathus rotulus (Alcock, 1898) *Deltocyathus sarsi (Gardiner & Waugh, 1938) *Deltocyathus stella Cairns & Zibrowius, 1997 *Deltocyathus suluensis Alcock, 1902 *Deltocyathus taiwanicus Hu, 1987 *Deltocyathus varians Gardiner & Waugh, 1938 *Deltocyathus vaughani Yabe & Eguchi, 1932 *Desmophyllum dianthus (Esper, 1794) 1 *Desmophyllum striatum Cairns, 1979 1 *Ericiocyathus echinatus Cairns & Zibrowius, 1997 Euphyllia ancora Veron & Pichon, 1979 Euphyllia cristata Chevalier, 1971 Euphyllia divisa Veron & Pichon, 1979 Euphyllia fimbriata (Spengler, 1799) Euphyllia glabrescens (Chamisso & Eysenhardt, 1821) Euphyllia paradivisa Veron, 1990 Euphyllia paraencora Veron, 1990 Euphyllia paraglabrescens Veron, 1990 Euphyllia yaeyamaensis (Shirai, 1980) Eusmilia fastigiata (Pallas, 1766) 1 *Goniocorella dumosa (Alcock, 1902) Gyrosmilia interrupta (Ehrenberg, 1834) +Heterocyathus aequicostatus M. Edwards & Haime, 1848 +Heterocyathus alternatus Verrill, 1865 *Heterocyathus hemisphericus Gray 1849 +Heterocyathus sulcatus (Verrill, 1866) *Hoplangia durotrix Gosse, 1860 *Labyrinthocyathus delicatus (Marenzeller, 1904) *Labyrinthocyathus facetus Cairns, 1979 1 *Labyrinthocyathus langae Cairns, 1979 1 *Labyrinthocyathus limatulus (Squires, 1964) *Labyrinthocyathus quaylei (Durham, 1947) *Lochmaeotrochus gardineri Cairns, 1999 *Lochmaeotrochus oculeus Alcock, 1902 *Lophelia pertusa (Linnaeus, 1758) 1 Montigyra kenti Matthai, 1928 Nemenzophyllia turbida Hodgson & Ross, 1981 *Nomlandia californica Durham & Barnard, 1952 *Oxysmilia circularis Cairns, 1998 *Oxysmilia corrugata Cairns, 1999 *Oxysmilia epithecata Cairns, 1999 *Oxysmilia rotundifolia (M. Edwards & Haime, 1848) ] *Paraconotrochus antarctica (Gardiner, 1929) *Paraconotrochus capense (Gardiner, 1904) *Paraconotrochus zeidleri Cairns & Parker, 1992 *Paracyathus anderssoni Duncan, 1899 *Paracyathus arcuatus Lindstrém, 1877 *Paracyathus cavatus Alcock, 1893 NO WWW WwW WH WwW WWW WwW WwW LoS) PHA HH HH HHH - = fhwohH aS *Paracyathus conceptus Gardiner & Waugh, 1938 *Paracyathus ebonensis Verrill, 1866 *Paracyathus fulvus Alcock, 1893 *Paracyathus humilis Verrill, 1870 *Paracyathus indicus indicus Duncan, 1889 *P. indicus gracilis Alcock, 1893 *Paracyathus lifuensis Gardiner, 1899 *Paracyathus molokensis Vaughan, 1907 *Paracyathus montereyensis Durham, 1947 *Paracyathus parvulus Gardiner, 1899 *Paracyathus porcellanus Verrill, 1866 *Paracyathus profundus Alcock, 1893 *Paracyathus pruinosus Alcock, 1902 *Paracyathus pulchellus (Philippi, 1842) *Paracyathus rotundatus Semper, 1872 *Paracyathus stearnsii Verrill, 1869 *Paracyathus stokesii M. Edwards & Haime, 1848 *Paracyathus vittatus Dennant, 1906 *Phacelocyathus flos (De Pourtalés, 1878) *Phyllangia americana americana M. Edwards & Haime, 1849 1 *P. americana mouchezii (Lacaze-Duthiers, 1897) *Phyllangia consagensis (Durham & Barnard, 1952) *Phyllangia dispersa Verrill, 1864 *Phyllangia echinosepes Ogawa, Takahashi & Sakai, 1997 *Phyllangia granulata Koch, 1886 *Phyllangia hayamaensis (Eguchi, 1968) *Phyllangia mouchezii (Lacaze-Duthiers, 1897) *Phyllangia papuensis Studer, 1878 Physogyra exerta Nemenzo & Ferraris, 1982 Physogyra lichtensteini (M. Edwards & Haime, 1851) Plerogyra eyrysepta Nemenzo, 1960 Plerogyra simplex Rehberg, 1892 Plerogyra sinuosa (Dana, 1846) Plerogyra turbida (Hodgson & Ross, 1981) *Polycyathus andamanicus Alcock, 1893 *Polycyathus atlanticus Duncan, 1876 *Polycyathus difficilis Duncan, 1876 *Polycyathus fulvus Wijsman-Best, 1970 *Polycyathus furanaensis Verheij & Best, 1987 *Polycyathus fuscomarginatus (Klunzinger, 1879) *Polycyathus hodgsoni Verheij & Best, 1987 *Polycyathus hondaensis (Durham & Barnard, 1952) *Polycyathus isabela Wells, 1982 *Polycyathus marigondoni Verheij & Best, 1987 *Polycyathus muellerae (Abel, 1959) *Polycyathus norfolkensis Cairns, 1995 *Polycyathus octuplus Cairns, 1999 *Polycyathus palifera (Verrill, 1869) *Polycyathus persicus (Duncan, 1876) *Polycyathus senegalensis Chevalier, 1966 *Polycyathus verrilli Duncan, 1889 NO N Ww BSS BRA HH HAHA 23 24 *Pourtalosmilia anthophyllites (Ellis & Solander, 1786) 2 *Pourtalosmilia conferta Cairns, 1978 *Premocyathus cornuformis (De Pourtalés, 1868) *Premocyathus dentiformis (Alcock, 1902) *Rhizosmilia gerdae Cairns, 1978 *Rhizosmilia elata Cairns & Zibrowius, 1997 *Rhizosmilia maculata (De Pourtalés, 1874) ] *Rhizosmilia multipaliferus Cairns, 1998 *Rhizosmilia robusta Cairns in Cairns & Keller, 1993 *Rhizosmilia sagamiensis (Eguchi, 1968) yes es NO *Solenosmilia variabilis Duncan, 1873 1 2. *Stephanocyathus campaniformis (Marenzeller, 1904) 2 *Stephanocyathus coronatus (De Pourtalés, 1867) 1 *Stephanocyathus crassus (Jourdan, 1895) p *Stephanocyathus diadema (Moseley, 1876) 1 *Stephanocyathus explanans (Marenzeller, 1904) *Stephanocyathus laevifundus Cairns, 1977 1 *Stephanocyathus moseleyanus (Sclater, 1886) 2 *Stephanocyathus nobilis (Moseley, 1873) | 2 *Stephanocyathus paliferus Cairns, 1977 l *Stephanocyathus platypus (Moseley, 1876) *Stephanocyathus regius Cairns & Zibrowius, 1997 *Stephanocyathus spiniger (Marenzeller, 1888) *Stephanocyathus weberianus Alcock, 1902 *Sympodangia albatrossi Cairns & Zibrowius, 1997 *Tethocyathus cylindraceus (De Pourtalés, 1868) 1 *Tethocyathus minor (Gardiner, 1899) *Tethocyathus recurvatus (De Pourtalés, 1878) 1 *Tethocyathus variabilis Cairns, 1979 1 a *Tethocyathus virgatus (Alcock, 1902) *Thalamophyllia gasti (Doderlein, 1913) 2 *Thalamophyllia gombergi Cairns, 1979 ] *Thalamophyllia riisei (Duchassaing & Michelotti, 1864) 1 *Thalamophyllia tenuescens (Gardiner, 1899) *Trochocyathus aithoseptatus Cairns, 1984 *Trochocyathus apertus Cairns & Zibrowius, 1997 *Trochocyathus brevispina Cairns & Zibrowius, 1997 *Trochocyathus burchae (Cairns, 1984) *Trochocyathus caryophylloides Alcock, 1902 *Trochocyathus cepulla Cairns, 1995 *Trochocyathus cinticulatus (Alcock, 1898) *Trochocyathus cooperi (Gardiner, 1905) *Trochocyathus decamera Cairns, 1994 *Trochocyathus discus Cairns & Zibrowius, 1997 *Trochocyathus efateensis Cairns, 1999 *Trochocyathus faciatus Cairns, 1979 1 *Trochocyathus fossulus Cairns, 1979 1 *Trochocyathus gardineri (Vaughan, 1907) *Trochocyathus gordoni Cairns, 1995 *Trochocyathus hastatus Bourne, 1903 *Trochocyathus japonicus Eguchi, 1968 RH HHH HH fA HHL *Trochocyathus longispina Cairns & Zibrowius, 1997 *Trochocyathus maculatus Cairns, 1995 *Trochocyathus mauiensis Vaughan, 1907 *Trochocyathus mediterraneus Zibrowius, 1980 *Trochocyathus oahensis Vaughan, 1907 *Trochocyathus patelliformis Cairns, 1999 *Trochocyathus philippinensis Semper, 1872 *Trochocyathus porphyreus (Alcock, 1893) *Trochocyathus rawsonii De Pourtalés, 1874 *Trochocyathus rhombocolumna Alcock, 1902 *Trochocyathus semperi Cairns & Zibrowius, 1997 *Trochocyathus spinosocostatus Zibrowius, 1980 *Trochocyathus vasiformis Bourne, 1903 *Vaughanella concinna Gravier, 1915 *Vaughanella margaritata (Jourdan, 1895) *Vaughanella multipalifera Cairns, 1995 *Vaughanella oreophila Keller, 1981 Dendrophylliidae *Astroides calycularis (Pallas, 1766) *Balanophyllia bairdiana M. Edwards & Haime, 1848 *Balanophyllia bayeri Cairns, 1979 *Balanophyllia bonaespei van der Horst, 1938 *Balanophyllia buccina Tenison-Woods, 1878 *Balanophyllia capensis Verrill, 1865 *Balanophyllia caribbeana Cairns, 1977 *Balanophyllia carinata (Semper, 1872) *Balanophyllia cedrosensis Durham, 1947 *Balanophyllia cellulosa Duncan, 1873 *Balanophyllia chnous Squires, 1962 *Balanophyllia corniculans Alcock, 1902 *Balanophyllia cornu Moseley, 1881 *Balanophyllia crassiseptum Cairns & Zibrowius, 1997 *Balanophyllia crassitheca Cairns, 1995 *Balanophyllia cumingii M. Edwards & Haime, 1848 *Balanophyllia cyathoides (De Pourtalés, 1871) *Balanophyllia dentata Tenison- Woods, 1879 *Balanophyllia desmophyllioides Vaughan, 1907 *Balanophyllia diademata van der Horst, 1927 *Balanophyllia diffusa Harrison & Poole, 1909 *Balanophyllia dineta Cairns, 1977 *Balanophyllia diomedeae Vaughan, 1907 *Balanophyllia dubia (Semper, 1872) *Balanophyllia elegans Verrill, 1864 *Balanophyllia elliptica (Tenison- Woods, 1878) *Balanophyllia elongata (Moseley, 1881) *Balanophyllia europaea (Risso, 1826) *Balanophyllia floridana De Pourtalés, 1868 *Balanophyllia galapagensis Vaughan, 1907 *Balanophyllia gemma (Moseley, 1881) *Balanophyllia gemmifera Klunzinger, 1879 NO eH HHA HA 25 26 *Balanophyllia generatrix Cairns & Zibrowius, 1997 *Balanophyllia gigas Moseley, 1881 *Balanophyllia hadros Cairns, 1979 *Balanophyllia imperialis Kent, 1871 *Balanophyllia iwayamaensis Abe, 1938 *Balanophyllia laysanensis Vaughan, 1907 *Balanophyllia malouiensis Squires, 1961 *Balanophyllia palifera De Pourtalés, 1878 *Balanophyllia parallela (Semper, 1872) *Balanophyllia parvula Moseley, 1881 *Balanophyllia pittieri Vaughan, 1919 *Balanophyllia ponderosa van der Horst, 1926 *Balanophyllia profundicella Gardiner, 1899 *Balanophyllia rediviva Moseley, 1881 *Balanophyllia regalis Alcock, 1893 *Balanophyllia regia Gosse, 1860 *Balanophyllia scabra Alcock, 1893 *Balanophyllia serrata Cairns & Zibrowius, 1997 *Balanophyllia stimpsonii Verrill, 1865 *Balanophyllia tenuis van der Horst, 1922 *Balanophyllia teres Cairns, 1994 *Balanophyllia thalassae Zibrowius, 1980 *Balanophyllia troprobanae Bourne, 1905 *Balanophyllia wellsi Cairns, 1977 *Balanophyllia yongei Crossland, 1952 *Bathypsammia falloscoialis Squires, 1959 *Bathypsammia tintinnabulum (De Pourtalés, 1868) *Cladopsammia echinata Cairns, 1984 *Cladopsammia eguchii (Wells, 1982) *Cladopsammia gracilis (M. Edwards & Haime, 1848) *Cladopsammia manuelensis (Chevalier, 1966) *Cladopsammia rolandi Lacaze-Duthiers, 1897 *Cladopsammia willeyi (Gardiner, 1900) *Dendrophyllia aculeata Latypov, 1990 *Dendrophyllia alcocki (Wells, 1954) *Dendrophyllia alternata De Pourtalés, 1880 *Dendrophyllia arbuscula van der Horst, 1922 *Dendrophyllia boschmai boschmai van der Horst, 1926 *D. boschmai cyathelioides Yabe & Eguchi, 1965 *Dendrophyllia californica Durham, 1947 *Dendrophyllia cladonia van der Horst, 1927 *Dendrophyllia cornigera (Lamarck, 1816) *Dendrophyllia cribrosa M. Edwards & Haime, 1851 *Dendrophyllia dilatata van der Horst, 1927 *Dendrophyllia florulenta Alcock, 1902 *Dendrophyllia ijimai Yabe & Eguchi, 1934 *Dendrophyllia incisa (Crossland, 1952) *Dendrophyllia indica Pillai, 1967 *Dendrophyllia johnsoni Cairns, 1991 *Dendrophyllia laboreli Zibrowius & Brito, 1984 *Dendrophyllia minuscula Bourne, 1905 NO N Hh _ = os *Dendrophyllia oldroydae Oldroyd, 1924 *Dendrophyllia ramea (Linnaeus, 1758) *Dendrophyllia robusta (Bourne, 1905) *Dendrophyllia velata Crossland, 1952 *Dichopsammia granulosa Song, 1994 Duncanopsammia axifuga (M. Edwards & Haime, 1848) *Eguchipsammia cornucopia (De Pourtalés, 1871) *Eguchipsammia fistula (Alcock, 1902) *Eguchipsammia gaditana (Duncan, 1873) *Eguchipsammia japonica (Rehberg, 1892) *Eguchipsammia serpentina (Vaughan, 1907) *Eguchipsammia wellsi (Eguchi, 1968) *Enallopsammia profunda (De Pourtalés, 1867) *Enallopsammia pusilla (Alcock, 1902) *Enallopsammia rostrata (De Pourtalés, 1878) *Endopachys bulbosa Cairns & Zibrowius, 1997 *Endopachys grayi M. Edwards & Haime, 1848 *Endopsammia philippensis M. Edwards & Haime, 1848 *Endopsammia pourtalesi (Durham & Barnard, 1952) *Endopsammia regularis (Gardiner, 1899) +Heteropsammia cochlea (Spengler, 1781) +Heteropsammia eupsammides (Gray, 1849) *Leptopsammia britannica (Duncan, 1870) *_eptopsammia chevalieri Zibrowius, 1980 *Leptopsammia columna Folkeson, 1919 *Leptopsammia crassa van der Horst, 1922 *Leptopsammia formosa (Gravier, 1915) *Leptopsammia poculum (Alcock, 1902) *Leptopsammia pruvoti Lacaze-Duthiers, 1897 *|_eptopsammia queenslandiae Wells, 1964 *Leptopsammia stokesiana M. Edwards & Haime, 1848 *Leptopsammia trinitatis Hubbard & Wells, 1987 *Notophyllia etheridgi Hoffmeister, 1933 *Notophyllia piscacauda Cairns, 1998 *Notophyllia recta Dennant, 1906 *Rhizopsammia annae (Van der Horst, 1933) *Rhizopsammia bermudensis Wells, 1972 *Rhizopsammia compacta Sheppard & Sheppard, 1991 *Rhizopsammia goesi (Lindstrém, 1877) *Rhizopsammia minuta van der Horst, 1922 *Rhizopsammia nuda van der Horst, 1926 *Rhizopsammia pulchra Verrill, 1870 *Rhizopsammia verrilli Van der Horst, 1922 *Rhizopsammia wellingtoni Wells, 1982 *Rhizopsammia wettsteini Scheer & Pillai, 1983 *Thecopsammia socialis De Pourtalés, 1868 *Trochopsammia infundibulum De Pourtalés, 1878 *Trochopsammia togata (Van der Horst, 1927) *Tubastraea coccinea Lesson, 1829 *Tubastraea diaphana (Dana, 1846) *Tubastraea faulkneri Wells, 1982 N NO WWW WwW WWW Ww Ww RA A HHH A Bs 27 Nn 28 *Tubastraea floreana Wells, 1982 *Tubastraea micranthus (Ehrenberg, 1834) *Tubastraea tagusensis Wells, 1982 Turbinaria bifrons Brueggemann, 1877 Turbinaria conspicua Bernard, 1896 Turbinaria crater (Pallas, 1766) Turbinaria frondens (Dana, 1846) Turbinaria heronensis Wells, 1958 Turbinaria irregularis Bernard, 1896 Turbinaria mesenterina (Lamarck, 1816) Turbinaria patula (Dana, 1846) Turbinaria peltata (Esper, 1794) Turbinaria radicalis Bernard, 1896 Turbinaria reniformis Bernard, 1896 Turbinaria stellulata (Lamarck, 1816) Faviidae Astreosmilia connata Ortmann, 1892 Australogyra zelli Veron, Pichon & Wijsman-Best, 1977 Barababattoia amicorum (M. Edwards & Haime, 1850) Barababattoia laddi (Wells, 1954) Barababattoia mirabilis Yabe & Sugiyama, 1941 Caulastrea curvata Wijsman-Best, 1972 Caulastrea echinulata (M. Edwards & Haime, 1849) Caulastrea furcata Dana, 1846 Caulastrea tumida Matthai, 1928 Colpophyllia amaranthus (O. F. Miiller, 1775) Colpophyllia breviserialis M. Edwards & Haime, 1849 Colpophyllia natans (Houttuyn, 1772) Cyphastrea agassizi (Vaughan, 1907) Cyphastrea chalcidicum (Forskal, 1775) Cyphastrea decadia Moll & Best, 1984 Cyphastrea japonica Yabe & Sugiyama, 1932 Cyphastrea microphthalma (Lamarck, 1816) Cyphastrea ocellina (Dana, 1846) Cyphastrea serailia (Forskal, 1775) Diploastrea heliopora (Lamarck, 1816) Diploria clivosa (Ellis & Solander, 1786) Diploria labyrinthiformis (Linnaeus, 1758) Diploria strigosa (Dana, 1846) Echinopora ashmorensis Veron, 1990 Echinopora forskaliana (M. Edwards & Haime, 1850) Echinopora fruticulosa Klunzinger, 1879 Echinopora gemmacea (Lamarck, 1816) Echinopora hirsutissima M. Edwards & Haime, 1849 Echinopora horrida Dana, 1846 Echinopora lamellosa (Esper, 1795) Echinopora mammiformis (Nemenzo, 1959) Echinopora pacificus Veron, 1990 Erythrastrea flabellata Pichon, Scheer & Pillai, 1983 Favia danae Verrill, 1872 WwW Ww WwW Ww Ww WWW W Ww Ww WwW ww « Ww Www Ww Ww WW WwW WH WwW = S$HHHHKHHA SHH HHA HHA HHH HK HA Favia favus (Forskal, 1775) Favia fragum (Esper, 1795) Favia gravida Verrill, 1868 Favia helianthoides Wells, 1954 Favia laxa (Klunzinger, 1879) Favia leptophylla Verrill, 1868 Favia lizardensis Veron, Pichon & Wijsman-Best, 1977 Favia maritima (Nemenzo, 1971) Favia matthaii Vaughan, 1918 Favia maxima Veron, Pichon & Wijsman-Best, 1977 Favia pallida (Dana, 1846) Favia rotumana (Gardiner, 1899) Favia rotundata (Veron, Pichon & Wijsman-Best, 1977) Favia speciosa (Dana, 1846) Favia stelligera (Dana, 1846) Favia veroni Moll & Best, 1984 Favia wisseli Scheer & Pillai, 1983 Favites abdita (Ellis & Solander, 1786) Favites chinensis (Verrill, 1866) Favites complanata (Ehrenberg, 1834) Favites flexuosa (Dana, 1846) Favites halicora (Ehrenberg, 1834) Favites pentagona (Esper, 1794) Favites peresi Faure & Pichon, 1978 Favites russelli (Wells, 1954) Favites stylifera Yabe & Sugiyama, 1937 Goniastrea aspera (Verrill, 1865) Goniastrea australensis (M. Edwards & Haime, 1857) Goniastrea deformis Veron, 1990 Goniastrea edwardsi Chevalier, 1971 Goniastrea favulus (Dana, 1846) Goniastrea palauensis (Yabe, Sugiyama & Eguchi, 1936) Goniastrea pectinata (Ehrenberg, 1834) Goniastrea retiformis (Lamarck, 1816) Leptastrea bewickensis Veron, Pichon & Wijsman-Best, 1977 Leptastrea bottae (M. Edwards & Haime, 1849) Leptastrea inaequalis Klunzinger, 1879 Leptastrea pruinosa Crossland, 1952 Leptastrea purpurea (Dana, 1846) Leptastrea transversa Klunzinger, 1879 Leptoria irregularis Veron, 1990 Leptoria phrygia (Ellis & Solander, 1786) Manicina areolata (Linnaeus, 1758) Montastraea annularis (Ellis & Solander, 1786) Montastraea annuligera (M. Edwards & Haime, 1849) Montastraea cavernosa Linnaeus, 1767 Montastraea curta (Dana, 1846) Montastraea faveolata (Ellis & Solander, 1786) Montastraea franksi (Gregory, 1895) Montastraea magnistellata Chevalier, 1971 Montastraea multipunctata Hodgson , 1985 Ww WwW WWW WWW WW WwW WWW WWW WW WwW Ww WW WWW WW WW WW WwW WwW dS pH HH HHA HAA aA HHH HHA HPA AHA HAHAH HHH HAHAHA 29 30 Montastraea valenciennesi (M. Edwards & Haime, 1848) Moseleya latistellata Quelch, 1884 Oulastrea crispata (Lamarck, 1816) Oulophyllia bennettae (Veron, Pichon & Best, 1977) Oulophyllia crispa (Lamarck, 1816) Parasimplastrea simplicitexta (Umbgrove, 1939) Platygyra contorta Veron, 1990 Platygyra crosslandi (Matthai, 1928) Platygyra daedalea (Ellis & Solander, 1786) Platygyra lamellina (Ehrenberg, 1834) Platygyra pini Chevalier, 1975 Platygyra ryukyuensis Yabe & Sugiyama, 1935 Platygyra sinensis (M. Edwards & Haime, 1849) Platygyra verweyi Wijsman-Best, 1976 Platygyra yaeyamaensis (Eguchi & Shirai, 1977) Plesiastrea versipora (Lamarck, 1816) Solenastrea bournoni M. Edwards & Haime, 1850 Solenastrea hyades (Dana, 1846) Flabellidae *Blastotrochus nutrix M. Edwards & Haime, 1848 *Falcatoflabellum rauolensis Cairns, 1995 *Flabellum alabastrum Moseley, 1876 *Flabellum angulare Moseley, 1876 *Flabellum angustum Yabe & Eguchi, 1942 *Flabellum aotearoa Squires, 1964 *Flabellum apertum apertum Moseley, 1876 *F. apertum borealis Cairns, 1994 *Flabellum arcuatile Cairns, 1999 *Flabellum areum Cairns, 1982 *Flabellum atlanticum Cairns, 1979 *Flabellum australe Moseley, 1881 *Flabellum campanulatum Holdsworth, 1862 *Flabellum chunii Marenzeller, 1904 *Flabellum conuis Moseley, 1881 *Flabellum curvatum Moseley, 1881 *Flabellum daphnense Durham & Barnard, 1952 *Flabellum deludens Marenzeller, 1904 *Flabellum flexuosum Cairns, 1982 *Flabellum floridanum Cairns, 1991 *Flabellum folkesoni Cairns, 1998 *Flabellum galapagense M. Edwards & Haime, 1848 *Flabellum gardineri Cairns, 1982 *Flabellum hoffmeisteri Cairns & Parker, 1992 *Flabellum impensum Squires, 1962 *Flabellum japonicum Moseley, 1881 *Flabellum knoxi Ralph & Squires, 1962 *Flabellum lamellulosum Alcock, 1902 *Flabellum lowekeyesi Squires & Ralph, 1965 *Flabellum macandrewi Gray, 1849 *Flabellum magnificum Marenzeller, 1904 i tO HHH HHH HHH HA HL Hh HWS *Flabellum marcus Keller, 1974 *Flabellum marenzelleri Cairns, 1989 *Flabellum messum Alcock, 1902 *Flabellum moseleyi De Pourtalés, 1880 1 *Flabellum ongulense Eguchi, 1965 *Flabellum patens Moseley, 1881 *Flabellum pavoninum Lesson, 1831 *Flabellum politum Cairns, 1989 *Flabellum sexcostatum Cairns, 1989 *Flabellum sibogae Gardiner, 1904 *Flabellum thouarsii M. Edwards & Haime, 1848 1 *Flabellum transversale transversale Moseley, 1881 *F. transversale conicum Yabe & Eguchi, 1942 *F. transversale triangulare Eguchi, 1965 *Flabellum tuthilli Hoffmeister, 1933 *Flabellum vaughani Cairns, 1984 *Javania antarctica (Gravier, 1914) *Javania borealis Cairns, 1994 *Javania cailleti (Duchassaing & Michelotti, 1864) 1 *Javania californica Cairns, 1994 *Javania exserta Cairns, 1999 *Javania fusca (Vaughan, 1907) *Javania insignis Duncan, 1876 *Javania lamprotichum Moseley, 1880 *Javania pseudoalabastra Zibrowius, 1974 1 *Monomyces pygmaea (Risso, 1826) *Monomyces rubrum (Quoy & Gaimard, 1833) *Placotrochides frustum Cairns, 1979 1 *Placotrochides scaphula Alcock, 1902 *Placotrochus laevis M. Edwards & Haime, 1848 *Placotrochus pedicellatus Tenison-Woods, 1879 *Polymyces fragilis (De Pourtalés, 1868) ] *Polymyces montereyensis (Durham, 1947) *Polymyces wellsi Cairns, 1991 1 *Rhizotrochus flabelliformis Cairns, 1989 *Rhizotrochus levidensis Gardiner, 1899 *Rhizotrochus niinoi Yabe & Eguchi, 1942 *Rhizotrochus tuberculatus (Tenison- Woods, 1879) *Rhizotrochus typus M. Edwards & Haime, 1848 *Truncatoflabellum aculeatum (M. Edwards & Haime, 1848) *Truncatoflabellum angiostomum (Folkeson, 1919) *Truncatoflabellum angustum Cairns & Zibrowius, 1997 *Truncatoflabellum arcuatum Cairns, 1995 *Truncatoflabellum australiensis Cairns, 1998 *Truncatoflabellum candeanum (M. Edwards & Haime, 1848) *Truncatoflabellum carinatum Cairns, 1989 *Truncatoflabellum crassum (M. Edwards & Haime, 1848) *Truncatoflabellum cumingii (M. Edwards & Haime, 1848) *Truncatoflabellum dens (Alcock, 1902) *Truncatoflabellum formosum Cairns, 1989 *Truncatoflabellum gardineri Cairns in Cairns & Keller, 1993 N WWW WwW vp HA BH HH & PPL HA RAP HHA HH Nn 31 32 *Truncatoflabellum inconstans (Marenzeller, 1904) *Truncatoflabellum incrustatum Cairns, 1989 *Truncatoflabellum irregulare (Semper, 1872) *Truncatoflabellum macroeschara Cairns, 1998 *Truncatoflabellum martensii (Studer, 1878) *Truncatoflabellum mortenseni Cairns & Zibrowius, 1997 *Truncatoflabellum multispinosum Cairns in Cairns & Keller, 1993 *Truncatoflabellum paripavoninum (Alcock, 1894) *Truncatoflabellum phoenix Cairns, 1995 *Truncatoflabellum pusillum Cairns, 1989 *Truncatoflabellum spheniscus (Dana, 1846) *Truncatoflabellum stabile (Marenzeller, 1904) 2 *Truncatoflabellum stokesi (M. Edwards & Haime, 1848) *Truncatoflabellum trapezoideum (Keller, 1981) *Truncatoflabellum truncum Cairns, 1982 *Truncatoflabellum vanuatu (Wells, 1984) *Truncatoflabellum veroni Cairns, 1998 *Truncatoflabellum vigintifarium Cairns, 1999 *Truncatoflabellum zuluense Cairns in Cairns & Keller, 1993 Fungiacyathidae *Fungiacyathus crispus (De Pourtalés, 187!) 1 2 *Fungiacyathus dennanti Cairns & Parker, 1992 *Fungiacyathus fissidiscus Cairns & Zibrowius, 1997 *Fungiacyathus fissilis Cairns, 1984 *Fungiacyathus fragilis Sars, 1872 1 2 *Fungiacyathus granulosus Cairns, 1989 *Fungiacyathus hydra Zibrowius & Gili, 1990 2 2 *Fungiacyathus marenzelleri (Vaughan, 1906) ] *Fungiacyathus margaretae Cairns, 1995 *Fungiacyathus multicarinatus Cairns, 1998 *Fungiacyathus paliferus (Alcock, 1902) *Fungiacyathus pliciseptus Keller, 1981 *Fungiacyathus pseudostephanus Keller, 1976 *Fungiacyathus pusillus pusillus (De Pourtalés, 1868) 1 *F. pusillus pacificus Cairns, 1995 *Fungiacyathus sandoi Cairns, 1999 *Fungiacyathus sibogae (Alcock, 1902) *Fungiacyathus stephanus (Alcock, 1893) *Fungiacyathus symmetricus (De Pourtalés, 1871) 1 *Fungiacyathus turbinolioides Cairns, 1989 *Fungiacyathus variegatus Cairns, 1989 Fungiidae Cantharellus doederleini (Marenzeller, 1907) Cantharellus jebbi Hoeksema, 1993 Cantharellus noumeae Hoeksema & Best, 1984 Ctenactis albitentaculata Hoeksema, 1989 Ctenactis crassa (Dana, 1846) Ctenactis echinata (Pallas), 1766 Fungia concinna Verrill, 1864 Ww WW Wd HHH ASH P= hh HKS ARS HA FH H HHH Fungia costulata Ortmann, 1889 Fungia curvata Hoeksema, 1989 Fungia cyclolites Lamarck, 1816 Fungia distorta Michelin, 1842 Fungia fragilis (Alcock, 1893) Fungia fralinae Nemenzo, 1955 Fungia fungites (Linnaeus, 1758) Fungia granulosa Klunzinger, 1879 Fungia gravis Nemenzo, 1955 Fungia hexagonalis M. Edwards & Haime, 1848 Fungia horrida Dana, 1846 Fungia moluccensis Van der Horst, 1919 Fungia paumotensis Stutchbury 1833 Fungia repanda Dana, 1846 Fungia scabra Déderlein, 1901 Fungia scruposa Klunzinger, 1879 Fungia scutaria Lamarck, 1801 Fungia seychellensis Hoeksema, 1993 Fungia sinensis (M. Edwards & Haime, 1851) Fungia somervillei Gardiner, 1909 Fungia spinifer Claereboudt & Hoeksema, 1987 Fungia taiwanensis Hoeksema & Dai, 1991 Fungia tenuis Dana, 1846 Fungia vaughani Boschma, 1923 Halomitra clavator Hoeksema, 1989 Halomitra pileus (Linnaeus, 1758) Heliofungia actiniformis (Quoy & Gaimard, 1833) Herpolitha limax (Esper, 1797) Lithophyllon mokai Hoeksema, 1989 Lithophyllon undulatum Rehberg, 1892 Podabacia crustacea (Pallas, 1766) Podabacia motuporensis Veron, 1990 Polyphyllia novaehiberniae (Lesson, 1831) Polyphyllia talpina (Lamarck, 1801) Sandalolitha dentata Quelch, 1884 Sandalolitha robusta (Quelch, 1886) Zoopilus echinatus Dana, 1846 Gardineriidae *Gardineria hawaiiensis Vaughan, 1907 *Gardineria minor Wells, 1973 1 *Gardineria paradoxa (De Pourtalés, 1868) 1 *Gardineria philippinensis Cairns, 1989 *Gardineria simplex (De Pourtalés, 1878) 1 Guyniidae *Guynia annulata Duncan, 1872 1 *Pedicellocyathus keyesi Cairns, 1995 *Pourtalocyathus hispidus (De Pourtalés, 1878) l *Schizocyathus fissilis De Pourtalés, 1874 1 *Stenocyathus vermiformis (De Pourtalés, 1868) 1 WW WW Ww WWW WW WWW WWW WW WwW WWW WWW WH WwW Ww WWW Ww FH HH AHHH HAHAHAHAHAHA pPHhHH HHP HHP HHP HP HPPHPHP HHA 33 34 *Temnotrochus kermadecensis Cairns, 1995 *Truncatoguynia irregularis Cairns, 1989 Meandrinidae Ctenella chagius Matthai, 1928 Dendrogyra cylindricus Ehrenberg, 1834 Dichocoenia stellaris M. Edwards & Haime, 1848 Dichocoenia stokesi M. Edwards & Haime, 1848 Meandrina meandrites (Linnaeus, 1758) Merulinidae Boninastrea boninensis Yabe & Sugiyama, 1935 Hydnophora bonsai Veron, 1990 Hydnophora exesa (Pallas, 1766) Hydnophora grandis Gardiner, 1906 Hydnophora microconos (Lamarck, 1816) Hydnophora pilosa Veron, 1985 Hydnophora rigida (Dana, 1846) Merulina ampliata (Ellis & Solander, 1786) Merulina scabricula Dana, 1846 Merulina scheeri Head, 1983 Paraclavarina triangularis (Veron, Pichon & Best, 1977) Scapophyllia cylindrica (M. Edwards & Haime, 1848) Micrabaciidae *Leptopenus antarcticus Cairns, 1989 *Leptopenus discus Moseley, 1881 *Leptopenus hypocoelus Moseley, 1881 *Leptopenus solidus Keller, 1977 *Letepsammia formosissima (Moseley, 1876) *Letepsammia fissilis Cairns, 1995 *Letepsammia franki Owens, 1994 *Letepsammia superstes (Ortmann, 1888) *Rhombopsammia niphada Owens, 1986 *Rhombopsammia squiresi Owens, 1986 *Stephanophyllia complicata Moseley, 1876 *Stephanophyllia fungulus Alcock, 1902 *Stephanophyllia neglecta Boschma, 1923 Mussidae Acanthastrea amakusensis Veron, 1990 Acanthastrea bowerbanki M. Edwards & Haime, 1857 Acanthastrea hemprichii (Ehrenberg, 1834) Acanthastrea echinata (Dana, 1846) Acanthastrea hillae Wells, 1955 Acanthastrea ishigakiensis Veron, 1990 Acanthastrea lordhowensis Veron & Pichon, 1982 Acanthastrea maxima Sheppard & Salm, 1988 Acanthastrea minuta Moll & Best, 1984 Acanthastrea rotundaflora Chevalier, 1975 eet WWW WWW WwW WH WwW WwW WWW WwW WwW oS pHtH HHH Hh p= Acanthophyllia deshayensiana (Michelin, 1850) Australomussa rowleyensis Veron, 1985 Blastomussa merleti (Wells, 1961) Blastomussa wellsi Wijsman-Best, 1973 Cynarina lacrymalis (M. Edwards & Haime, 1848) Indophyllia macassarensis Best & Hoeksema, 1987 Isophyllastrea rigida (Dana, 1846) ] Isophyllia sinuosa (Ellis & Solander, 1786) 1 Lobophyllia corymbosa (Forskal, 1775) Lobophyllia costata (Dana, 1846) Lobophyllia diminuta Veron, 1985 Lobophyllia hataii Yabe, Sugiyama & Eguchi, 1936 Lobophyllia hemprichii (Ehrenberg, 1834) Lobophyllia pachysepta Chevalier, 1975 Lobophyllia robusta Yabe, Sugiyama & Eguchi, 193 Mussa angulosa (Pallas, 1766) Mussismilia braziliensis (Verrill, 1868) Mussismilia harttii (Verrill, 1868) Mussismilia hispida (Verrill, 1901) Mycetophyllia aliciae Wells, 1973 Mycetophyllia daniana M. Edwards & Haime, 1849 Mycetophyllia ferox Wells, 1973 Mycetophyllia lamarckiana M.Edwards &Haime, 1848 Mycetophyllia reesi Wells, 1973 Scolymia australis (M. Edwards & Haime, 1849) Scolymia cubensis M. Edwards & Haime, 1849 1 Scolymia lacera (Pallas, 1766) 1 Scolymia vitiensis Brueggemann, 1877 Scolymia wellsii Laborel, 1967 1 Symphyllia agaricia M. Edwards & Haime, 1849 Symphyllia erythraea (Klunzinger, 1879) Symphyllia hassi Pillai & Scheer, 1976 Symphyllia radians M. Edwards & Haime, 1849 Symphyllia recta (Dana, 1846) Symphyllia valenciennesii M. Edwards & Haime, 1849 Symphyllia wilsoni Veron, 1985 Oculinidae Acrhelia horrescens (Dana, 1846) *Archohelia rediviva Wells & Alderslade, 1979 *Bathelia candida Moseley, 1881 l *Cyathelia axillaris (Ellis & Solander, 1786) Galaxea alta Nemenzo, 1980 Galaxea astreata (Lamarck, 1816) Galaxea fascicularis (Linnaeus, 1767) Galaxea paucisepta Claereboudt, 1990 *Madrepora arbuscula (Moseley, 1881) *Madrepora carolina (De Pourtalés, 1871) 1 *Madrepora kauaiensis Vaughan, 1907 *Madrepora minutiseptum Cairns & Zibrowius, 1997 *Madrepora oculata Linnaeus, 1758 1 WW WW Ww WWW WW Ww WWW WW Ww Ww SA HHH SH HHH HAH PS ah HHH 35 36 *Madrepora porcellana Moseley, 1881 Oculina arbuscula L. Agassiz, 1864 +Oculina diffusa Lamarck, 1816 Oculina patagonica De Angelis, 1908 *Oculina profunda Cairns, 1991 Oculina robusta De Pourtalés, 1871 +Oculina tenella De Pourtalés, 1871 Oculina valenciennesi M. Edwards & Haime, 1850 +Oculina varicosa Lesueur, 1821 *Oculina virgosa Squires, 1958 Schizoculina fissipara (M. Edwards & Haime, 1850) *Sclerhelia hirtella (Pallas, 1766) Simplastrea vesicularis Umbgrove, 1940 Pectiniidae Echinophyllia aspera (Ellis & Solander, 1786) Echinophyllia echinata (Saville-Kent, 1871) Echinophyllia echinoporoides Veron & Pichon, 1979 Echinophyllia maxima Moll & Best, 1984 Echinophyllia nishihirai Veron, 1990 Echinophyllia orpheensis Veron & Pichon, 1979 Echinophyllia patula (Hodgson & Ross, 1981) Echinophyllia tosaensis Yabe & Eguchi, 1935 Mycedium elephantotus (Pallas, 1766) Mycedium robokaki Moll & Best, 1984 Oxypora crassispinosa Nemenzo, 1980 Oxypora glabra Nemenzo, 1959 Oxypora lacera (Verrill, 1864) Pectinia alcicornis (Saville-Kent, 1871) Pectinia elongata Rehberg, 1892 Pectinia lactuca (Pallas, 1766) Pectinia paeonia (Dana, 1846) Pectinia teres Nemenzo, 1981 Physophyllia ayleni (Wells, 1934) Pocilloporidae *Madracis asanoi Yabe & Sugiyama, 1936 +Madracis asperula M. Edwards & Haime, 1849 *Madracis brueggemanni (Ridley, 1881) Madracis decactis (Lyman, 1859) Madracis formosa Wells, 1973 *Madracis hellana M. Edwards & Haime, 1850 *Madracis interjecta Marenzeller, 1907 *Madracis kauaiensis Vaughan, 1907 Madracis kirbyi Veron & Pichon, 1976 Madracis mirabilis sensu Wells, 1973 *Madracis myriaster (M. Edwards & Haime, 1849) +Madracis pharensis (Heller, 1868) *Madracis profunda Zibrowius, 1980 Madracis senaria Wells, 1974 *Madracis singularis Rehberg, 1892 SS ee el WWW Ww WwW WW Ww Ww Wn WW AHHH RHA HAHHH HHH HHHAHA Palauastrea ramosa Yabe & Sugiyama, 1941 Pocillopora capitata Verrill, 1864 Pocillopora damicornis (Linnaeus, 1758) Pocillopora elegans Dana, 1846 Pocillopora eydouxi M. Edwards & Haime, 1860 Pocillopora meandrina Dana, 1846 Pocillopora verrucosa (Ellis & Solander, 1786) Pocillopora woodjonesi Vaughan, 1918 Seriatopora caliendrum Ehrenberg, 1834 Seriatopora hystrix Dana, 1846 Stylophora kuehImanni Scheer & Pillai, 1983 Stylophora mamillata Scheer & Pillai, 1983 Stylophora mordax (Dana, 1846) Stylophora pistillata (Esper, 1797) Stylophora wellsi Scheer, 1964 Poritidae Alveopora allingi Hoffmeister, 1925 Alveopora catalai Wells, 1968 Alveopora excelsa Verrill, 1864 Alveopora fenestrata (Lamarck, 1816) Alveopora gigas Veron, 1985 Alveopora japonica Eguchi, 1968 Alveopora marionensis Veron & Pichon, 1982 Alveopora ocellata Wells, 1954 Alveopora spongiosa Dana, 1846 Alveopora tizardi Bassett-Smith, 1890 Alveopora verrilliana Dana, 1872 Alveopora viridis (Quoy & Gaimard, 1833) Goniopora burgosi Nemenzo, 1955 Goniopora cellulosa Veron, 1990 Goniopora columna Dana, 1846 Goniopora djiboutiensis Vaughan, 1907 Goniopora eclipsensis Veron & Pichon, 1982 Goniopora fruticosa Saville-Kent, 1891 Goniopora lobata M. Edwards & Haime, 1860 Goniopora minor Crossland, 1952 Goniopora norfolkensis Veron & Pichon, 1982 Goniopora palmensis Veron & Pichon, 1982 Goniopora pandoraensis Veron & Pichon, 1982 Goniopora pendulus Veron, 1985 Goniopora planulata (Ehrenberg, 1834) Goniopora polyformis Zou, 1980 Goniopora savignyi Dana, 1846 Goniopora somaliensis Vaughan, 1907 Goniopora stokesi M. Edwards & Haime, 1851 Goniopora stutchburyi Wells, 1955 Goniopora tenella (Quelch, 1886) Goniopora tenuidens Quelch, 1886 Porites annae Crossland, 1952 Porites aranetai Nemenzo, 1955 WWW WWW WW WW WwW WWW WwW Ww WWW WwW Ww WWW WWW WH WWW WWW WH WwW WWW WwW WH WwW Ww Ww pRHHHH HA AHH ABRDLA RADA HAAAHAHAAADAAAAHAHAAAHAAHA ppp HR HHH Ann nnnn 317 38 Porites astreoides Lamarck, 1816 Porites attenuata Nemenzo, 1955 Porites australiensis Vaughan, 1918 Porites baueri Squires, 1959 Porites branneri Rathbun, 1888 Porites colonensis Zlatarski, 1990 Porites compressa Dana, 1846 Porites cumulatus Nemenzo, 1955 Porites cylindrica Dana, 1846 Porites deformis Nemenzo, 1955 Porites densa Vaughan, 1918 Porites echinulata Klunzinger, 1879 Porites eridani Umbgrove, 1940 Porites evermanni Vaughan, 1907 Porites furcata Lamarck, 1816 Porites gabonensis Gravier, 1911 Porites heronensis Veron, 1985 Porites horizontalata Hoffmeister, 1925 Porites iwayamaensis Eguchi, 1938 Porites latistella Quelch, 1884 Porites lichen Dana, 1846 Porites lobata Dana, 1846 Porites lutea M. Edwards & Haime, 1860 Porites mayeri Vaughan, 1918 Porites myrmidonensis Veron, 1985 Porites negrosensis Veron, 1990 Porites nigrescens Dana, 1846 Porites nodifera Klunzinger, 1879 Porites okinawensis Veron, 1990 Porites panamensis Verrill, 1866 Porites porites (Pallas, 1766) Porites rus (Forskal, 1775) Porites sillimaniani Nemenzo, 1976 Porites solida (Forskal, 1775) Porites somaliensis Gravier, 1910 Porites stephensoni Crossland, 1952 Porites sverdrupi Durham, 1947 Porites undulata (Klunzinger, 1879) Porites vaughani Crossland, 1952 Stylaraea punctata (Linnaeus, 1758) Rhizangiidae * Astrangia atrata (Dennant, 1906) *Astrangia browni Palmer, 1928 * Astrangia californica Durham & Barnard, 1952 *Astrangia conferta Verrill, 1870 * Astrangia costata Verrill, 1866 * Astrangia dentata Verrill, 1866 * Astrangia equatorialis Durham & Barnard, 1952 * Astrangia haimei Verrill, 1866 * Astrangia howardi Durham & Barnard, 1952 No Ww Ww WWW W WwW Ww WwW WwW AHH HH & 4 fPHpHphHhtppeeHf Nn AANA AaaAnn *Astrangia macrodentata Theil, 1940 *Astrangia mercatoris Theil, 1941 +Astrangia poculata (Ellis & Solander, 1786) *Astrangia rathbuni Vaughan, 1906 *Astrangia solitaria (Lesueur, 1817) *Astrangia woodsi Wells, 1955 *Cladangia exusta Ltitken, 1873 *Cladangia gemmans Chevalier, 1966 *Culicia australiensis Hoffmeister, 1933 *Culicia cuticulata Klunzinger, 1879 *Culicia excavata M. Edwards & Haime, 1849 *Culicia fragilis Chevalier, 1971 *Culicia hoffmeisteri Squires, 1966 *Culicia quinaria Tenison-Woods, 1878 *Culicia rubeola (Quoy & Gaimard, 1833) *Culicia smithii (M. Edwards & Haime, 1849) *Culicia stellata Dana, 1848 *Culicia subaustraliensis Ogawa, Takahashi & Sakai, 1997 *Culicia tenella tenella Dana, 1848 *C. tenella natalensis (Duncan, 1876) *Culicia tenuisepes Ogawa, Takahashi & Sakai, 1997 *Culicia verreauxi M. Edwards & Haime, 1850 *Oulangia bradleyi Verrill, 1866 *Oulangia cyathiformis Chevalier, 1971 *Oulangia stokesiana stokesiana M. Edwards & Haime, 1848 *O. stokesiana miltoni Yabe & Eguchi, 1932 Siderastreidae Anomastrea irregularis Marenzeller, 1901 Coscinaraea columna (Dana, 1846) Coscinaraea crassa Veron & Pichon, 1980 Coscinaraea exaesa (Dana, 1846) Coscinaraea fossata (Dana, 1846) Coscinaraea hazimanensis Yabe & Sugiyama, 1936 Coscinaraea marshae Wells, 1962 Coscinaraea mcneilli Wells, 1962 Coscinaraea monile (Forskal, 1775) Coscinaraea wellsi Veron & Pichon, 1980 Horastrea indica Pichon, 1971 Psammocora brighami Vaughan, 1907 Psammocora contigua (Esper, 1797) Psammocora digitata M. Edwards & Haime, 1851 Psammocora explanulata Van der Horst, 1922 Psammocora haimeana M. Edwards & Haime, 1851 Psammocora nierstraszi Van der Horst, 1921 Psammocora obtusangula (Lamarck, 1816 ) Psammocora profundacella Gardiner, 1898 Psammocora stellata Verrill, 1866 Psammocora superficialis Gardiner, 1898 Psammocora vaughani Yabe & Sugiyama, 1936 Pseudosiderastrea tayamai Yabe & Sugiyama, 1935 NO WWW WH WwW WWW WWW WWW WH WwW Ww HH HHL RAH HHH SHH H 39 40 Siderastrea glynni Budd & Guzman, 1994 Siderastrea radians (Pallas, 1766) 1 Siderastrea savignyana M. Edwards & Haime, 1850 Siderastrea siderea (Ellis & Solander, 1786) 1 Trachyphylliidae Trachyphyllia geoffroyi (Audouin, 1826) Turbinoliidae *Alatotrochus rubescens (Moseley, 1876) * Australocyathus vincentinus (Dennant, 1904) *Conocyathus gracilis Cairns, 1998 *Conocyathus zelandiae Duncan, 1876 *Cryptotrochus brevipalus Cairns, 1999 *Cryptotrochus carolinensis Cairns, 1988 1 *Cryptotrochus javanus Cairns, 1988 *Cyathotrochus herdmani Bourne, 1905 *Cyathotrochus nascornatus Gardiner & Waugh, 1938 *Cyathotrochus pileus (Alcock, 1902) *Deltocyathoides orientalis (Duncan, 1876) *Deltocyathoides stimpsonii (De Pourtalés, 1871) l *Dunocyathus parasiticus Tenison- Woods, 1878 *Endocyathopora laticostata Cairns, 1989 *Foveolocyathus alternans (Cairns & Parker, 1992) *Foveolocyathus verconis Dennant, 1904 *Holcotrochus crenulatus Dennant, 1904 *Holcotrochus scriptus Dennant, 1902 *Idiotrochus emarciatus Duncan, 1865 *Idiotrochus kikutii (Yabe & Eguchi, 1941) *Kionotrochus suteri Dennant, 1906 *Notocyathus conicus (Alcock, 1902) *Notocyathus venustus (Alcock, 1902) *Peponocyathus dawsoni Cairns, 1995 *Peponocyathus folliculus (De Pourtalés, 1868) | *Peponocyathus minimus (Yabe & Eguchi, 1937) *Platytrochus compressus (Tenison-Woods, 1878) *Platytrochus hastatus Dennant, 1902 *Platytrochus laevigatus Cairns & Parker, 1992 *Platytrochus parisepta Cairns & Parker, 1992 *Pleotrochus venustus (Alcock, 1902) *Pleotrochus zibrowii Cairns, 1997 *Pseudocyathoceras avis (Durham & Barnard, 1952) *Sphenotrochus andrewianus M. Edwards & Haime, 1848 *Sphenotrochus aurantiacus Marenzeller, 1904 *Sphenotrochus auritus De Pourtalés, 1874 1 *Sphenotrochus evexicostatus Cairns in Cairns & Keller, 1993 *Sphenotrochus excavatus Tenison-Woods, 1878 *Sphenotrochus gardineri Squires, 1961 *Sphenotrochus gilchristi Gardiner, 1904 *Sphenotrochus hancocki Durham & Barnard, 1952 *Sphenotrochus imbricaticostatus Cairns in Cairns & Keller, 1993 in) WWW WwW WwW WwW Ww oa WwW Ww SHH H HH HHA *Sphenotrochus ralphae Squires, 1964 *Sphenotrochus squiresi Cairns, 1995 *Thrypticotrochus multilobatus Cairns, 1989 *Thrypticotrochus petterdi (Dennant, 1906) *Trematotrochus corbicula (De Pourtalés, 1878) *Trematotrochus hedleyi Dennant, 1906 *Tropidocyathus labidus Cairns & Zibrowius, 1997 *Tropidocyathus lessoni (Michelin, 1842) *Turbinolia stephensoni (Wells, 1959) Incertae sedis *Cylicia inflata De Pourtalés, 1878 Milleporidae Millepora alcicornis Linnaeus, 1758 Millepora boschmai De Weerdt & Glynn, 1991 Millepora braziliensis Verrill, 1868 Millepora complanata Lamarck, 1816 Millepora dichotoma (Forskal, 1775) Millepora exaesa (Forskal, 1775) Millepora foveolata Crossland, 1952 Millepora intricata Edwards, 1857 Millepora latifolia Boschma, 1948 Millepora murrayi Quelch, 1884 Millepora nitida Verrill, 1868 Class Hydrozoa Order Capitata Millepora platyphylla Hemprich & Ehrenberg, 1834 Millepora squarrosa Lamarck, 1816 Millepora striata Duchassaing & Michelotti, 1864 Millepora tenera Boschma, 1949 Millepora tuberosa Boschma, 1966 Millepora xishaensis Zou, 1978 *Hydractiniidae Hydrocorella africana Stechow, 1921 Janaria mirablis Stechow, 1921 Polyhydra calcarea (Carter, 1877) *Stylasteridae Adelopora crassilabrum Cairns, 1991 Adelopora fragilis Cairns, 1991 Adelopora moseleyi Cairns, 1991 Adelopora pseudothyron Cairns, 1982 Astya aspidopora Cairns, 1991 Astya subviridis (Moseley, 1879) —_ Order Filifera NO SHH BBR APH HHA 41 42 Calyptopora reticulata Boschma, 1968 Calyptopora sinuosa Cairns, 1991 Cheiloporidion pulvinatum Cairns, 1983 1 Conopora adeta Cairns, 1987 Conopora anthohelia Cairns, 1991 Conopora candelabrum Cairns, 1991 Conopora dura Hickson & England, 1909 Conopora gigantea Cairns, 1991 Conopora laevis (Studer, 1878) Conopora tetrastichopora Cairns, 1991 Conopora unifacialis Cairns, 1991 Conopora verrucosa (Studer, 1878) Crypthelia affinis Moseley, 1879 Crypthelia balia Hickson & England, 1905 Crypthelia clausa Broch, 1947 Crypthelia cryptotrema Zibrowius, 1981 Crypthelia curvata Cairns, 1991 Crypthelia cymas Cairns, 1986 Crypthelia dactylopoma Cairns, 1986 Crypthelia eueides Cairns, 1986 Crypthelia floridana Cairns, 1986 | Crypthelia formosa Cairns, 1983 Crypthelia fragilis Cairns, 1983 Crypthelia gigantea Fisher, 1938 Crypthelia glebulenta Cairns, 1986 Crypthelia glossopoma Cairns, 1986 | Crypthelia insolita Cairns, 1986 1 Crypthelia japonica (M. Edwards & Haime, 1849) Crypthelia lacunosa Cairns, 1986 Crypthelia medioatlantica Zibrowius & Cairns, 1992 Crypthelia micropoma Cairns, 1985 Crypthelia papillosa Cairns, 1986 1 Crypthelia peircei De Pourtalés, 1867 ] Crypthelia platypoma Hickson & England, 1905 Crypthelia polypoma Cairns, 1991 Crypthelia pudica M. Edwards & Haime, 1849 Crypthelia ramosa Hickson & England, 1905 Crypthelia robusta Cairns, 1991 Crypthelia stenopoma Hickson & England, 1905 Crypthelia studeri Cairns, 1991 Crypthelia tenuiseptata Cairns, 1986 1 Crypthelia trophostega Fisher, 1938 Crypthelia vascomarquesi Zibrowius & Cairns, 1992 Cyclohelia lamellata Cairns, 1991 Distichopora anceps Cairns, 1978 Distichopora anomala Cairns, 1986 1 Distichopora barbadensis De Pourtalés, 1874 1 Distichopora borealis borealis Fisher, 1938 D. borealis japonica Broch, 1942 Distichopora cervina De Pourtalés, 1871 1 Distichopora coccinea Gray, 1860 i) a HHH HL P= HHH HHH nN fon) Distichopora contorta De Pourtalés, 1878 1 Distichopora dispar Cairns, 1991 Distichopora foliacea De Pourtalés, 1868 1 Distichopora gracilis Dana, 1848 Distichopora irregularis Moseley, 1879 Distichopora laevigranulosa Cairns, 1986 Distichopora livida Tenison-Woods, 1879 Distichopora nitida Verrill, 1864 Distichopora profunda Hickson & England, 1909 Distichopora providentiae Hickson & England, 1909) Distichopora rosalindae Cairns, 1986 1 Distichopora serpens Broch, 1942 Distichopora sulcata De Pourtalés, 1867 1 Distichopora uniserialis Cairns, 1986 1 Distichopora vervoorti Cairns & Hoeksema, 1999 Distichopora violacea (Pallas, 1766) Distichopora yucatanensis Cairns, 1986 1 Errina altispina Cairns, 1986 1 Errina antarctica (Gray, 1872) Errina aspera (Linnaeus, 1767) Errina atlantica Hickson, 1912 Errina bicolor Cairns, 1991 Errina boschmai Cairns, 1983 Errina capensis Hickson, 1912 Errina chathamensis Cairns, 1991 Errina cheilopora Cairns, 1983 Errina cochleata Pourtalés, 1867 1 Errina cooki Hickson, 1912 Errina cyclopora Cairns, 1983 Errina dabneyi (De Pourtalés, 1871) Errina dendyi Hickson, 1912 Errina fissurata Gray, 1872 Errina gracilis Marenzeller, 1903 1 Errina hicksoni Cairns, 1991 Errina japonica Eguchi, 1968 Errina kerguelensis Cairns, 1983 Errina laevigata Cairns, 1991 Errina laterorifa Eguchi, 1964 Errina macrogastra Marenzeller, 1904 Errina novaezelandiae Hickson, 1912 Errina porifera Naumov, 1960 Errina reticulata Cairns, 1991 Errina sinuosa Cairns, 1991 Errinopora cestoporina Cairns, 1983 Errinopora latifundata Naumov, 1960 Errinopora nanneca Fisher, 1938 Errinopora pourtalesi (Dall, 1884) Errinopora stylifera (Broch, 1935) Errinopora zarhyncha Fisher, 1938 Errinopsis fenestrata Cairns, 1983 Errinopsis reticulum Broch, 1951 BSS & pS PS Shhh Aaan fon) lon 43 44 Gyropora africana Boschma, 1960 Inferiolabiata labiata (Moseley, 1879) Inferiolabiata lowei (Cairns, 1983) Inferiolabiata spinosa Cairns, 1991 Lepidopora acrolophos Cairns, 1983 Lepidopora biserialis Cairns, 1986 Lepidopora carinata (De Pourtalés, 1867) Lepidopora clavigera Cairns, 1986 Lepidopora concatenata Cairns, 1991 Lepidopora cryptocymas Cairns, 1985 Lepidopora decipiens Boschma, 1964 Lepidopora dendrostylus Cairns, 1991 Lepidopora diffusa Boschma, 1963 Lepidopora eburnea (Calvet, 1903) Lepidopora glabra (De Pourtalés, 1867) Lepidopora granulosa Cairns, 1983 Lepidopora microstylus Cairns, 1991 Lepidopora polystichopora Cairns, 1985 Lepidopora sarmentosa (Boschma, 1968) Lepidopora symmetrica Cairns, 1991 Lepidotheca altispina Cairns, 1991 Lepidotheca brochi Cairns, 1986 Lepidotheca cervicornis (Broch, 1942) Lepidotheca chauliostylus Cairns, 1991 Lepidotheca fascicularis (Cairns, 1983) Lepidotheca horrida (Hickson & England, 1905) Lepidotheca inconsuta Cairns, 1991 Lepidotheca macropora Cairns, 1986 Lepidotheca pourtalesi Cairns, 1986 Lepidotheca ramosa (Hickson & England, 1905) Lepidotheca robusta Cairns, 1991 Lepidotheca tenuistylus (Broch, 1942) Paraerrina decipiens Broch, 1942 Phalangopora regularis Kirkpatrick, 1897 Pliobothrus echinatus Cairns, 1986 Pliobothrus fistulosus Cairns, 1991 Pliobothrus gracilis Zibrowius & Cairns, 1992 Pliobothrus symmetricus De Pourtalés, 1868 Pliobothrus tubulatus (De Pourtalés, 1867) Pseudocrypthelia pachypoma (Hickson & England, 1905) Sporadopora dichotoma (Moseley, 1877) Sporadopora micropora Cairns, 1991 Sporadopora mortenseni Broch, 1942 Stellapora echinata (Moseley, 1879) Stenohelia concinna Boschma, 1964 Stenohelia conferta Boschma, 1968 Stenohelia echinata Eguchi, 1968 Stenohelia maderensis (Johnson, 1862) Stenohelia pauciseptata Cairns, 1986 Stenohelia profunda Moseley, 1881 Stenohelia tiliata (Hickson & England, 1905) NN a Stenohelia umbonata (Hickson & England, 1905) Stenohelia yabei (Eguchi, 1941) Stephanohelia praecipua Cairns, 1991 Stylantheca papillosa (Dall, 1884) Stylantheca petrograpta (Fisher, 1938) Stylantheca porphyra Fisher, 1931 Stylaster alaskanus Fisher, 1938 Stylaster amphiheloides Kent, 1871 2 Stylaster antillarum Zibrowius & Cairns, 1982 1 Stylaster asper Kent, 1871 Stylaster aurantiacus Cairns, 1986 1 Stylaster bellus (Dana, 1848) Stylaster bilobatus Hickson & England, 1905 Stylaster bithalamus Broch, 1936 2 Stylaster blatteus (Boschma, 1961) 2 Stylaster bocki Broch, 1936 Stylaster boreopacificus Broch, 1932 Stylaster boschmai (Eguchi, 1965) Stylaster brochi (Fisher, 1938) Stylaster brunneus Boschma, 1970 Stylaster californicus (Verrill, 1866) Stylaster campylecus campylecus (Fisher, 1938) S. campylecus parageus (Fisher, 1938) S. campylecus tylotus (Fisher, 1938) S. campylecus trachystomus (Fisher, 1938) Stylaster cancellatus Fisher, 1938 Stylaster carinatus Broch, 1936 Stylaster cocosensis Cairns, 1991 Stylaster complanatus De Pourtalés, 1867 1 Stylaster corallium Cairns, 1986 1 Stylaster crassior Broch, 1936 Stylaster densicaulis Moseley, 1879 1 Stylaster dentatus Broch, 1936 Stylaster divergens Marenzeller, 1904 Stylaster duchassaingi De Pourtalés, 1867 1 Stylaster eguchii (Boschma, 1966) Stylaster elassotomus Fisher, 1938 Stylaster erubescens erubescens De Pourtalés, 1868 1 S. erubescens groenlandicus Zibrowius & Cairns, 1992 S. erubescens britannicus Zibrowius & Cairns, 1992 S. erubescens meteorensis Zibrowius & Cairns, 1992 Stylaster eximius Kent, 1871 Stylaster filogranus De Pourtalés, 1871 1 Stylaster flabelliformis (Lamarck, 1816) Stylaster galapagensis Cairns, 1986 Stylaster gemmascens (Esper, 1794) 2 Stylaster gracilis M. Edwards & Haime, 1850 Stylaster granulosus M. Edwards & Haime, 1850 Stylaster hattorrii (Eguchi, 1968) Stylaster horologium Cairns, 1991 Stylaster ibericus Zibrowius & Cairns, 1992 NNN WN i) AA HAHA SHH LH nmAnnn NAanrnnn 45 46 Stylaster imbricatus Cairns, 1991 Stylaster incompletus (Tenison-Woods, 1883) Stylaster incrassitus (Eguchi, 1941) Stylaster inornatus Cairns, 1986 l Stylaster laevigatus Cairns, 1986 1 Stylaster lonchitis Broch, 1947 Stylaster marenzelleri Cairns, 1986 Stylaster maroccanus Zibrowius & Cairns, 1992 Stylaster marshae Cairns, 1988 Stylaster microstriatus Broch, 1936 Stylaster miniatus (De Pourtalés, 1868) 1 Stylaster moseleyanus (Fisher, 1938) Stylaster multiplex Hickson & England, 1905 Stylaster nobilis (Kent, 1871) Stylaster norvegicus (Gunnerus, 1768) Stylaster papuensis Zibrowius, 1981 Stylaster polymorphus Broch, 1936 Stylaster polyorchis (Fisher, 1938) Stylaster profundus (Moseley, 1879) ] Stylaster profundiporus Broch, 1936 Stylaster pulcher Quelch, 1884 Stylaster purpuratus (Naumov, 1960) Stylaster ramosus Broch, 1947 Stylaster robustus (Cairns, 1983) Stylaster rosaceus (Greeff, 1886) Stylaster roseus (Pallas, 1766) ] Stylaster sanguineus Valenciennes in M. Edw. & Haime,1850 Stylaster scabiosus Broch, 1935 Stylaster solidus Broch, 1935 Stylaster spatula Cairns, 1986 ] Stylaster stejnegeri (Fisher, 1938) Stylaster stellulatus Stewart, 1878 Stylaster subviolacea (Kent, 1871) Stylaster tenisonwoodsi Cairns, 1988 Stylaster venustus (Verrill, 1870) Stylaster verrillii (Dall, 1884) Systemapora ornata Cairns, 1991 tO & 4 Unknown a PS ATOLL RESEARCH BULLETIN NO. 460 ATOLLS AS SETTLEMENT LANDSCAPES: UJAE, MARSHALL ISLANDS BY MARSHALL I. WEISLER ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. AUGUST 1999 165° 170° ® Pikaar Qnewetak —Pikinni So = QRondik . Aelsaina Fo" Toke ™ Utrok . - Aelok = Enylamieg Swe 08a ilo) ° Likiep Q& “Jomo %, Wojja Kwajalein**, ayers . ve [X] eLae =) %+ SAdkup_Maloelap Biginnigar a OS Sa) oem “> Q ie Namo Aur ° @ Jowat 4 =< 2, Dacisriaplap Q “+ Maye AEs Sr R e Kill MARSHALL ISLANDS Namdik@ 5° @Ebon 170° S Ebbetyu Bikku NQ ¢ Rua a Wotya © Langeba Bee , Bokerok i) FS 4 & oy Bo 165° 40' FIGURE 1. Ujae Atoll and archaeological sites with map of the Marshall Islands showing location of Ujae. Atolls in bold (Utrok, Maloelap and Ebon), situated along the continuum of rainfall from the dry north to the wet south, are part of the long-term archaeological study of the Marshall Islands. ATOLLS AS SETTLEMENT LANDSCAPES: UJAE, MARSHALL ISLANDS BY MARSHALL I. WEISLER! ABSTRACT Williamson and Sabath (1982) have demonstrated a significant relationship between modern population size and environment by examining atoll area and rainfall in the Marshall Islands. The present work seeks to extend that argument into prehistory by examining the relationship of ancient habitation sites and size of aroid pit agricultural systems to atoll land area and rainfall regime along the 1,500-3,500 mm precipitation gradient in the Marshall Islands. Four atolls were selected for study: Ebon at the wettest extent in the extreme south; Ujae and Maloelap near the center of the archipelago; and Utrok at the dry north. The first phase of this long-term archaeological program is reported. During the survey of Ujae Atoll (9° 05'N, 165° 40' E), three habitation sites, an aroid pit agricultural zone, one early historic burial, and seven fish traps, weirs, and enclosures were recorded. Along with excavations at two habitation sites (8 m? total area), 35 traditional artifacts were recovered (shell adzes, ornaments, and manufacturing tools). Seven radiocarbon age determinations document land use beginning as early as the third century A.D. A beachrock sample dated to 2450 + 70 BP relates to atoll development. Some 4,748 bones of fish, birds, turtles, Pacific rats, lizards, humans, and possible cetaceans, along with nearly 13 kg of shellfish, provide the basis for understanding prehistoric subsistence, human adaptations to the atoll setting, and land use patterns. INTRODUCTION Pacific coral atolls are unquestionably the most precarious landscapes for settlement, yet many of them evidence continuous human occupation for 2,000 years. Unlike the high volcanic islands of the Pacific plate and the non-oceanic or continental land masses west of the Andesite Line, coral atolls are unique in their small size, low elevation, limited diversity of terrestrial flora and fauna, poorly developed soils, and an absence of surface potable water—all characteristics that would limit a sustained human presence. How did small human founding groups survive and, in a sense, flourish on these most challenging of Pa- cific landscapes? In 1993, a long-term interdisciplinary study of the archaeology and pre- 'Department of Anthropology, University of Otago, P. O. Box 56, Dunedin, New Zealand. Email: marshall.weisler@ stonebow.otago.ac.nz Manuscript received 9 October 1998; revised 1 April 1999 2 history of the Marshall Islands was initiated. Located about 4,000 km southwest of Hawai‘i, the Marshalls are situated mostly between 4° - 12° north latitude and consist of 29 low coral atolls and five small islands oriented in two roughly parallel lines trending southeast-north- west for about 1,100 km. Selection of four atolls for intensive archaeological survey and excavations was based on the marked rainfall gradient beginning in the wet south, with about 3,500 mm of annual precipitation, to 1,500 mm in the dry north (Fig. 1). On Pacific atolls, rainfall is the most significant determinant of species diversity (Fosberg 1984; Stoddart 1992) and is mirrored in the prehistoric settlement patterns by varying densities of surface artifacts, area of prehistoric habitations, and size of agricultural systems. Williams and Sabath (1982), in their modern population studies, have demonstrated the close correspond- ence between human carrying capacity, island size, and climate. Consequently, Marshall Islands atolls were selected for study at opposite ends of the rainfall gradient beginning with the most southern one of Ebon (4° N latitude), Maloelap (8° 50') and Ujae (9°) near the center of the archipelago, and Utrok (11° 15’), one of the farthest north, permanently inhab- ited atolls. Beginning about 2000 BP, human colonists of the Marshalls targeted the pristine stocks of fish, sea birds, turtles, and coconut crabs (Birgus latro), yet permanent settlers would ultimately depend on terrestrial production for the bulk of their subsistence. Aroid pit culti- vation for Giant swamp taro (Cyrtosperma chamissonis)—uniquely adapted to the harsh conditions of atolls (Thompson 1982; Weisler in press)—is the foundation of Marshall Islands culture. Pandanus, with numerous varieties cultivated throughout the Marshalls (Stone 1960), was the fundamental tree crop for the dry northern atolls, while breadfruit increased in importance towards the south. Arrowroot (Jacca leontopetaloides) was grown on all the atolls and supplemented the carbohydrate intake. After a brief review of previous archaeological research in the Marshalls, results from the first archaeological survey and excavations from this long-term project are reported. PREVIOUS ARCHAEOLOGICAL STUDIES Archaeological research in the Marshall Islands began with the Kelton-Bishop Museum Expedition in 1977, and continued with phases in 1979 and 1980. The first effort was a brief reconnaissance-level survey of 12 atolls and recording of 42 sites with the goal of determining if these atolls had sufficient, intact prehistoric cultural deposits warranting detailed study (Rosendahl 1987). More than 4,000 portable artifacts were collected and included shell adzes, fishing gear, scrapers, ornaments, and manufacturing tools such as shell hammerstones and pumice and coral abraders. Although not from dated contexts, this is an important archaeological collection of Marshall Islands artifacts. Eight habitation sites on four atolls were test excavated and one radiocarbon age determination established occu- pation by 1260 + 80 BP (Rosendahl 1987:161). While Rosendahl’s survey was extensive, Riley, in 1979, focussed on a single atoll, completing an intensive survey and transect excavations of a major prehistoric village site. He recorded 122 sites on Majuro Atoll which he classified into midden areas (sites with surface food remains), house platforms, coral-faced structures (some of which were his- toric burials, e.g., Riley [1987:fig. 2.4]), fishtraps, wells, and a Cyrtosperma pit zone. Fol- 3 lowing the observations made by the botanist Hatheway (1953), Riley suggested that the oldest portions of habitation sites should be in the middle of the islet. The greatest density of recorded sites was at Laura village, on the largest islet, and excavation conducted there produced a radiocarbon age determination of 1970 + 110 BP—one of the oldest habitation dates for the Marshalls. The last phase of the Kelton-Bishop Expedition was a detailed survey of Arno Atoll with systematic transect excavations undertaken to locate buried prehistoric sites that may not have surface indications (Dye 1987). Some 133 islets were surveyed and 164 sites recorded, mostly habitations. The earliest human occupation was dated to about 1000 BP, while dates for the reef platform were 2,500-3,000 years old. In the northern Marshalls, Streck surveyed and conducted test excavations on five atolls (1990). Brief excavations on Bikini Atoll yielded 35 radiocarbon age determinations, 15 percent of which were older than 2000 BP. Most were “grab” samples from eroded expo- sures of cultural layers and the stratigraphic details have not been published. Few archae- ologists accept the oldest dates because old drift logs may have been used for fuel (Kirch and Weisler 1994:292). That is, dating drift wood from trees with a long life span can add hundreds of years to radiocarbon age determinations. Consequently, the date when the log was used for fuel is increased by the age of the tree. In conjunction with expansion of the airport on Kwajalein Atoll, Shun and Athens (1990; see also Beardsley 1994) reported a buried gray layer on Kwajalein islet dated to about 2000 BP. Located near the center of the islet, the layer may represent a natural swamp or constructed aroid pit. Also on Kwajalein, Weisler et al. (in press) recovered a single human burial during construction of utility lines. Found with 151 grave goods, the bones probably are those of an individual of relatively high status who may have shared communication or trade links with groups on nearby Polynesian islands. Ancestral ties with the Marianas are suggested by a comparison of ancient mtDNA. Widdicombe recently completed a technological study of Marshallese shell adzes from Ebon, Maloelap, and Ujae atolls (1997). Adzes were manufactured from large and small taxa of Tridacna, helmet shells (Cassidae), conches (Lambis sp.), and less frequently from cones (Conidae) and augers (Terebra sp.). No significant differences were noted in terms of assemblage composition between Maloelap and Ebon yet, today, residents of the largest islet of Kaven, Maloelap Atoll, say that T: gigas does not grow there—perhaps because of the deep offshore lagoon waters. Consequently, 7. gigas adzes may have been imported. Somerville-Ryan (1998) has reported on a taphonomic investigation of archaeological shell- fish assemblages from Ebon Atoll. During prehistory, shellfish were brought to sites as food, as raw material for tool and ornament manufacture, and in water-rolled coral gravel used for pavement. His study suggests that only a few taxa reliably reflect food use, and archaeologists should be aware of this when selecting specimens for radiocarbon dating. The past two decades produced at least brief archaeological surveys on nearly half of the Marshall Islands, and a range of prehistoric artifacts and habitation sites have been documented. Detailed excavations are, as yet, rare, but widely accepted radiocarbon dates establish human occupation by about 2000 BP. ARCHAEOLOGICAL SURVEY OF UJAE ATOLL Introduction and Objectives The main focus of the archaeological study was the survey of Ujae islet and subsequent transect excavations for defining the site boundaries—at least the ocean to lagoon extent. This was done for three reasons: (1) the most substantial and oldest village is normally located on the largest islet of an atoll; (2) the entire population of the atoll now lives on Ujae islet, which facilitated the work and made best use of the limited time; and (3) the archaeo- logical study was also undertaken as part of a historic preservation training program and, by conducting most work in the village, it was easier to interact with school children and other members of the community who visited the on-going excavations. In my former capacity as Chief Archaeologist for the Republic of the Marshall Islands, I found that it is much easier to teach historic preservation by involving residents directly in archaeological survey and excavations, than by lecturing about abstract concepts in the village schoolhouse. Located near the center of the archipelago in the Ralik chain of the Marshalls, Ujae Atoll (9° 05' N, 165° 40' E)—ranks 22 in land area amongst the group—consists of 12 larger islets (with a total land area of 1.8 km?) which surround a 180 km? lagoon (Bryan 1971). The largest islet, named for the atoll, contains the entire population of 488 people (Office of Planning and Statistics 1988). Some 81% of households are involved in copra production and credit is issued for kerosene, rice, flour, sugar, and cooking oil. It is a sub- sistence lifestyle with nearly all of the people engaged in growing food and fishing. In 1988, firewood was the dominant source of cooking fuel, but less so in 1994 after several years of U. S. Federal food assistance. Only one diesel boat was operating at the time of our visit and it was rented for $US25/ day including two operators. The archaeological survey began after we left Ujae islet mid- morning on 4 July 1994 traveling within the lagoon up the windward side stopping at Langeba, Wotya, bypassing Rua, and spending the night at Ebbetyu (Fig. 1). Leaving Ebbetyu the following morning, we bypassed the small islets of Bikku and Erlie and had to wait several hours for high tide before approaching Enylamieg. We only spent a short time there to enable us to reach Bokerok with sufficient time to set up camp before dark. Bokerok supports a dense colony of nesting noddies and I saw it as a good opportunity to record how the Ujae Marshallese capture, butcher, and cook birds, and to document the fracture and burning characteristics these activities have on the bones (see Weisler and Gargett 1993) as well as the spatial patterns resulting from discard. This ethnoarchaeological study was of value for interpreting bird bones recovered from the excavations on Ujae and elsewhere. We spent two nights on Bokerok before going to Bock where we stayed an additional two nights while excavating a midden site. We returned to Ujae by mid-day on the 9th. The results of the survey are presented below by islet, beginning with Ujae and following coun- ter-clockwise around the lagoon. Islet Survey Ujae Islet (0.47 km). This is the largest islet of the atoll and the one suspected of having the biggest and oldest village. The center of the islet has numerous elongate and sinuous Cyrtosperma pits which today are mostly abandoned. Breadfruit, banana, papaya, and FIGURE 2. Aerial view of Ujae taken 17 February 1944 before construction of the landing strip now located along the southern half of the islet. The modern village paths are clearly visible just inland of the lagoon shore. Note the extensive reef flats that cover most of this frame. (Courtesy of Bernice P. Bishop Museum, negative No. CP 117,087, composite of #116 and #117, AP28A-VD3.) 6 Pandanus trees are maintained in the modern village which is located within the northern third of the islet. In Figure 2, an aerial photograph taken 17 February 1944, the main village paths form an obtuse triangle just inland from the lagoon shore near the middle of the coast, with one leg of the polygon extending to the northwest point. The first step in the archaeo- logical survey of Ujae islet was to walk around the modern village and observe the stratigraphy in trash pits and foundations dug for new houses. This was an expedient way of gaining a general indication of the depth and location of prehistoric deposits. In gardening areas with clear ground visibility, large, fragmented shellfish (such as Tridacna and Lambis) were observed and a few shell adzes collected. Several transects walked through the inte- rior established that the ancient village (designated site 1) was located lagoonward of the Cyrtosperma pits. Seven 1 m? units were excavated along a transect beginning near the center of the islet to just above the lagoon shore; an additional unit was excavated 250 m east of the main transect. The lagoon-ocean side transect established that the site was 420m wide (north-south) and a conservative estimate (based on the distribution of surface arti- facts and cultural layers seen in modern trash pits) suggest a site length of ca. 600 m along the lagoon shore axis. A site area of 25 hectares seems reasonable. On the reef flat off the west coast, six stone configurations, presumably fish traps, weirs, and enclosures, were located. No Ujae residents knew anything about these features since they probably went out of use decades before; these features are described below. Langeba Islet (0.06 km7). This is the first islet north of Ujae along the east or windward margin of the atoll. No archaeological sites were located on this islet as it is small with numerous wind-fallen trees, and wave-deposited rubbish—evidence of wash-overs. There are less than 100 coconut trees on the islet, with typical coastal vegetation such as Tournefortia and Scaevola. Coconut crabs were collected from the bases of a few large Pisonia grandis trees in the interior of the islet. A large exposure of beachrock was noted along southeast coast. The gravelly interior and small islet size does not support a fresh water lens. Coconut crabs found today, in addition to nesting seabirds in previous times, would have been the dominant food resources and the main reasons for visiting the islet. Wotya Islet (0.16 km?). Nearly in the center of the east coast of the atoll, Wotya may be large enough to support a fresh water lens which is suggested by the modern habitation area. Turtle nests were observed on the ocean side, north shore. Arrowroot was seen grow- ing in the interior and large Pisonia trees that cover the south quarter of the islet support nesting noddies (Anous spp.). The presence of the small tree, Neisosperma sp. (Ochrosia sp.), in the islet interior, suggested to Fosberg (1955:21) that storm waves occasionally reached inland areas, dispersing seeds of this plant, which otherwise should not have been found growing away from the shoreline. A prehistoric midden (site 2) was located along the lagoon side west shore immediately above an area of beachrock. The shoreline is eroded here, forming a wave-cut bank ca. 1.25 m high exposing a dark midden layer ca. 0.5 m thick consisting of Tridacna and Lambis shellfish, coral oven stones, and one earth oven, the base of which was 1.10 m below surface. The site runs about 75 m along the shoreline, while its inland extent can only be determined by excavation. However, it is probably less than 0.5 hectare. Numerous Tridacna valves inland of the exposure may be from the modern settle- ment. 7 Rua Islet (0.08 km?). We sailed slowly past this islet and due to its small size, low eleva- tion, and sparse vegetation, its was unlikely to have prehistoric cultural deposits. Ebbetyu Islet (0.22 km”). We arrived at this islet after 5 pm and could only spend a short time surveying before dark. Turtle tracks were noted on the northwest point on the ocean side. Two modern, abandoned houses were located just inland of the south shore on the lagoon side. The most southeast house had three large breadfruit trees and a well about 50 m inland from a coral pavement. The islet size suggests a prehistoric site should be present, but excavation would be necessary to locate buried deposits. Compounding the problem of identifying prehistoric cultural material on the ground surface is the presence of large shell- fish that may only be refuse from the modern occupants; this is especially true of histori- cally-discarded shellfish that have been bleaching in the sun for many decades and, conse- quently, can be confused with prehistoric midden. Fosberg noted, in February 1952, a layer of pumice pebbles covering a few square yards of Anuij, the islet south and connected to Ebbetyu (Fosberg 1955:20). Pumice is used occasionally today to sharpen tools and would have been an important resource in prehistory. Bikku Islet (0.09 km?). We did not have time to stop at this islet but, as we sailed by, it was clear from its small size, rocky topography, and sparse vegetation that no prehistoric cul- tural deposits would be present. Erlie Islet (0.06 km”). One of the smallest of the islets, time limits did not permit stopping at this locale but, similar to the environmental constraints observed on Bikku, no prehis- toric cultural deposits could be expected. Fosberg reported a pure Cordia forest which he considered a rare occurrence in the Marshalls (1955:22). He mentioned an “enormous number of hermit crabs” which, today, are the preferred bait for bottom fishing for the lethrinid Monotaxis grandoculis (personal observation on Utrok Atoll, 1996-7). Biginnigar Islet (0.07 km). This small shoal could not have supported prehistoric cultural deposits. Residents mentioned that Golden cowries (Cypraea aurantium) are occasionally collected on the adjacent reef flats. Enylamieg Islet (0.44 km). This islet must be approached at high tide to affect a landing. Unfortunately, we arrived at low tide and wasted precious time walking across the broad reef flat. Arriving at the southern tip of the islet, we walked inland about one-half of the way, noting a village (abandoned perhaps a few generations ago) marked by beachrock slab-lined paths near the west coast. Because this islet is no longer used for habitation, it is densely overgrown and ground visibility was negligible. No prehistoric sites or features were observed, but islet size and presence of the historic village suggests that an ancient site must surely exist. Bock Islet (0.09 km”). On the leeward side of Ujae Atoll, just north of the main channel to the lagoon, Bock islet has a dense, black, prehistoric midden located just above a beachrock formation on the lagoon shore. Fosberg and colleagues established an astronomical station on Bock for several days in early 1952. On the south end of the islet, large Pisonia trees reached heights of 30 m—a fact remarked by Fosberg in 1952 (1955:19). He also recorded 8 large individuals of Intsia, at least 25 m in height, which, in Hawai‘i, is normally a “small to medium-sized tree” (Neal 1965:418). The economically important Morinda citrifolia (nen) was recorded in several places along an elevation transect. The temporary encampment, noted by Fosberg some 30 years before, was still in use in 1994; and the Marshallese are still coming to Bock to capture the plentiful noddies. A partially submerged beachrock “island” is the best marker for the prehistoric site, where, beginning just above the beach and immediately west of the site, a trail leads inland to a modern sleeping house, cook shed, and well. The inland extent of the midden is clearly delimited by a wet swale less than 10 m distant from a low mound, while the site measures 135 m east along the lagoon shore (see Figs 1 and 3). Assuming a width of ca. 15 m, the site is about 2,025 m? (or 1/5th hectare). Unlike the largest islets of most atolls, this smaller landmass does not have a central depres- sion where Cyrtosperma pits are usually found. On Bock, the islet rises gradually from the lagoon shore before descending to the cobble ridge and beachrock on the ocean side (Fig. 3). We stayed on the islet a day and a half completing an elevation transect and a 1 m? excavation unit, described below. Bock Islet, Site 3 Elevation transect few coconut, dense cobbles and boulders,| dense coconut, rolling and uneven ground surface Pandanus, Pisonia grandis, sandy-gravel ground fo} a oO » = ij Excavation unit 1 tural deposit cobble ridge cobble and sand beach oO > a oO o 3 beach rock vertical 5x FIGURE 3. Transect through site 3, Bock Island. Bokerok Islet (<0.05 km7). This small islet does not have prehistoric cultural deposits but, at least today, is a significant nesting area for noddies. On the lagoon side there is a small camp used for a few days at a time while capturing noddies. The site lies about 4 m inland of the vegetation line and consists of a level area (9 m*) used for sleeping as well as process- ing and roasting birds. A fire pit, used for roasting and not baking birds, is situated near the level area and about | m away is a dense concentration of wing bones where these non- edible portions of birds were routinely discarded. Elements containing high meat portions were tossed around the perimeter of the sleeping area after roasting and consumption (Fig. 4). Two historic burials were found inland, one was mapped and consisted of a rectangular perimeter of beachrock slabs, only two remaining in the original upright position (Fig. 5). These are typical historic burials as described by Finsch during his visit to the Marshalls in the 1870’s (Finsch 1893). Bokerok Islet Ujae Atoll | beach 8m Approximate North FIGURE 4. Modern bird processing camp, Bokerok islet. Bokerok Islet, Site 4, Historic Burial Z Upright beachrock slab Fallen beachrock slab FIGURE 5. Historic burial, site 4, Bock islet. 10 Site Classes Habitation Sites. The common habitation site was a midden recognized on the surface by scattered and fragmented large shellfish (such as Tridacna and Lambis), coral and beachrock oven stones, and shell adzes. Earth ovens are commonly used on the outer atolls (not Majuro or Kwajalein) and shellfish are still eaten today. Consequently, it is not always possible to differentiate old, bleached, and weathered food shells that are, say 100 years old, from prehistoric specimens. Therefore, the presence of shell adzes was the best criterion to iden- tify a prehistoric habitation site, especially in reference to topographic setting and proxim- ity to aroid pits. Buried prehistoric cultural deposits were recognized by dark, sandy sedi- ment, earth ovens, and dispersed charcoal, oven stones, shell artifacts, shellfish, and bone midden. Cultural deposits were found exposed at the lagoon shoreline near beachrock (e.g., site 2) or seen in the walls of trash pits and other recent excavations. There has been a tendency by some archaeologists to designate a new site for each seemingly isolated sur- face scatter of artifacts and shellfish found on an islet as Rosendahl did for portions of Ebon islet, Ebon Atoll (1987:26, 84). However, even on large islets, excavation has demonstrated that there is one main prehistoric village site associated with inland aroid pits. Predictably, the major habitation site was located on the largest islet of Ujae associated with aroid pits for the cultivation of Giant swamp taro. Estimated to be at least 25 hectares, it is larger than all the other habitation sites on the atoll combined. Site 2 on Wotya, and site 3 on Bock are, together, less than a hectare and are not associated with aroid pits. Agricultural Sites. Aroid pits are typically, but not exclusively, located near the interior of the largest islets where the Ghyben-Herzberg fresh water lens is the thickest and most reli- able during periods of drought (Weisler 1999). Extensive aroid pits were observed in the interior of Ujae islet, but similar features should be present on Enylamieg as well. The Ujae pits were mostly long and sinuous forms, meandering for 100 m or more. Three pits ranged in width from 28 to 35 m, and nearly 2 m deep as measured from the rims. Only one small pit (20 by 34 m) was planted with Cyrtosperma (see Fig. 6), while a portion of a much larger pit was currently being weeded and planted. Because most of the pits are no longer used and, consequently, overgrown with dense, high vegetation, the time-consuming task of clearing these features to make accurate size assessments was not undertaken. These features were not recorded in detail as the main emphasis of the archaeological study and training program was on habitation sites. However, an elevation transect on Ujae islet es- tablished the location of the aroid pit zone relative to the ancient village site and shorelines (Fig. 7). Fish Traps. Situated close to the lagoon shore on the reef flat are various stacked stone configurations, some of which are clearly fish traps, but other features may have had differ- ent functions such as enclosures for holding fish or turtles. These devices show signs of disrepair and probably have not been used for decades. Riley (1987:187) provided a classi- fication of fish traps from Majuro Atoll but, unfortunately, did not provide illustrations of his types. I adapt his scheme (using his type descriptions) in the presentation of the Ujae features. Riley’s type | is most common on Majuro and Ujae as well. It consists of a V- shaped configuration that funnels fish to a circular trap (Fig. 8A). Most Ujae examples are situated open to the ocean side and two of the traps (site 5B and C) take advantage of the 1] FIGURE 6. Small Cyrtosperma pit with a few arrowroot plants in foreground and to right. Ujae Islet, Site 1 Elevation transect | Cyrtosperma pit zone | modern village ocean few coconut, dense cobbles lagoon and boulders, rolling and uneven ground surface ® ® 2) O © © A 2540 + 70 BP Pot Ct Cl ii if OW > Sey iyos) CerOmecec ONIN GUpet cobble ridge ch rock sea level sea level ° a 3 vertical 10x = So So So o Oi Oy wow Ok NO Oa ii) GH fish bone [__) shellfish urchin pumice FIGURE 7. Elevation transect through site 1, Ujae islet. Fish bone (counts) and shellfish, urchins, and pumice weights for excavated units. shoreline which acts as an extension of the landward leg of trap site SC (Fig. 8). Site 6 is an adaptation of type | traps that takes advantage of the lagoonal topography, such as natural alignments of coral, rubble, and depressions, as well as the coastline for extending the 12 FIGURE 8. Aerial photo of Ujae islet, site 5 fish traps features A-C, possible turtle enclosure feature D, wall feature E, and enclosure feature F. Photo taken 7 February 1944. (Bernice P. Bishop Museum negative number CP 117,081.) length of the catchment (Fig. 9). This trap is four times as long and three times as wide as the west coast features of similar form. On Majuro, large schools of the Bigeye scad (Selar crumenopthalmus) are caught in similar traps (personal communication, Laura, Majuro resident, 1994). The Yellowstripe goatfish (Mulloides flavolineatus and M. vanicolensis), which congregates in large schools over sand flats, may also be a species commonly caught in stone traps. Site 5E uses the local topography to aid in directing fish—a similar situation to the type 1 traps. Extending from the shoreline, a low wall was built perpendicular to a ridge on the reef flat, thus forming a three-sided square open to the ocean side (Figs 8 and 10, middle). There is no circular trap as in the type | features, while this square configuration may have been used in conjunction with a seine net that closed the fourth side of the trap. The two legs extending from site 5F may have acted to help channel fish to the square enclosure, but the exact function of this feature is uncertain (Fig. 10, lower). It is somewhat similar to the description of Riley’s type 3 traps (1987:187). Site SD is a circular enclosure measuring about 3 m in diameter (Fig. 10, upper). This may have been a turtle pen as similar, but larger, structures have been noted elsewhere in the Pacific on islands with shal- 13 FIGURE 9. Aerial photo of Ujae islet, site 6 fish trap. Photo taken 17 February 1944. (Bernice P. Bishop Museum negative number CP 117,086.) low offshore topography (Emory 1939:17). Riley’s type 2, a semicircle that joins at the beach, was not seen on Ujae Atoll. In summary, Ujae fish traps were built to take advantage of the local offshore topogra- phy and were situated in reference to prevailing currents. Both topography and currents acted to enhance the efficiency of the traps, which could be checked after each incoming tide. Although a range of fish species were undoubtedly trapped, Bigeye scad and goatfish may have accounted for significant catches. I have observed seine net catches of hundreds of rabbitfishes (Siganidae) on Utrok Atoll, and it is reasonable to suspect that species of this family may have been caught in traps as well. The variability of forms in Ujae fish traps and enclosures warrants a more inclusive classificatory scheme for the Marshall Islands fea- tures that should be adapted from Riley’s initial attempt at describing the Majuro sites. It is worth noting that traps were only associated with Ujae islet probably reflecting the rela- tively larger village there. These devices may be a form of intensification of the marine subsistence regime, built during late prehistory. It is unlikely that traps would have been constructed as part of the colonization phase of settlement when marine stocks were at their highest levels and capturing fish required less effort. 14 Se eee | FIGURE 10. Site 5 possible turtle enclosure feature D (top), wall feature E (middle) built perpendicular to natural reef alignment, and enclosure feature F (bottom). i) TEST EXCAVATIONS Excavations were conducted primarily on Ujae islet, where 8 m* were completed at the major habitation (site 1), with an additional 1 m? dug at site 3 on Bock islet. The objectives were to: (1) determine the depth and nature of the prehistoric cultural deposits; (2) obtain samples for radiocarbon dating; (3) collect shellfish and faunal material to address subsist- ence issues; (4) acquire artifacts from secure stratigraphic contexts; and, in the case of site 1, (5) determine the lagoon to ocean side site boundaries. Experience has shown that it is simply not possible to accurately determine habitation site boundaries on atolls without excavation. Unit excavations proceeded by arbitrary 10 cm levels or spits within, but never crossing, stratigraphic boundaries. All sediments were passed through 1/4" (6.4 mm) sieves, then material sorted into one of several classes which normally included: artifacts, bone, shell, urchins, crustacea, and charcoal. Separating cultural material into these classes reduced the amount of breakage during shipment and facilitated lab work. The stratigraphic profile of at least one side of each excavation unit was drawn, recording such layer characteristics as boundary, Munsell color, texture, consistency, structure, etc. Elevation transects were made through the excavation units from the lagoon shore to the ocean side showing the topo- graphic features and vegetation of the islet. Site MLUj-1, Ujae Islet At the major habitation site of Ujae Atoll, seven one meter square units were excavated along a north-south transect running due south (180°) for 420 m from the middle of the lagoon shore (Fig. 7). As time permitted, an eighth unit was placed 250 m east of unit 2 to determine the limits of the site boundary along that axis. Stratigraphic profiles are pre- sented in Figures 11 and 12 and detailed layer descriptions provided in Appendix 1. The cultural layer was thickest and oldest on the lagoon side, just north of the aroid pits, and generally thinned north and south of this point. In general, the upper 20 cm of each excava- tion unit contained dense coconut roots in a gravelly sand matrix. In the midst of the mod- ern village, unit 5 had a surface layer of water-rolled coral pebbles—a typical pavement found throughout villages today. Except in units 2 and 5, all cultural material was contained in the top 25 cm in a gravelly sand to coarse sand, black to dark gray matrix. Portions of combustion features were encountered and included earth ovens (see Fig. 11, unit 5, S profile) and other smaller and thinner features that may have been ovens, or charcoal and oven stone concentrations may have served for roasting food. The culturally sterile subsoil was encountered near 75 cm below surface and, in unit 2, below one meter. Cultural content is described below. Site MLUj-3, Bock Islet This small midden site, located just inland of a beachrock formation, runs parallel to the lagoon shore. The site does not extend inland very far, yet is quite long, suggesting that inhabitants preferred to be close to the water, away from the rocky interior. One | m2 unit was excavated at the highest point of the site to get a complete sample of the stratigraphy. Five layers were defined primarily on the basis of color, texture, and artifact content (see Fig. 13 and Appendix 1). Dense gravel was encountered to one meter below surface sug- 16 Site 1 Unit 1, E profile fine sand dense gravel coarse sand @ coarse sand @) Site 1 Unit 3, E profile gravel pavement coconut root ‘ Site 1 Unit 5, E profile charcoal Site 1 Unit 2, N profile gravelly sand gravelly sand Beta-74845 1660 + 60 BP —> grades to sterile Site 1 Unit 4, W profile Site 1 Unit 5, S profile Slightly hard sand dense gravel grades to coarse sand __ loose, coarse sand Beta-76018 560 +70 BP oven stones Ein Cee oe ee ee FIGURE 11. Profiles of units 1-5, site 1, Ujae islet. IW, Site 1 Site 1 Unit 6, W profile looseveand Unit 7, W profile coconut root 2 weakly coherent sand Beta-79582 $8.0. ¢ medium sand 101.3 0.7% . Beta-79581 } area of dispersed charcoal 140996 +0.6% _—-_so— mottled sand coarse sand Site 1 Unit 8, E profile dense gravel grades to coarse sand sterile sand ©) FIGURE 12. Profiles of units 6-8, site 1, Ujae islet. gesting a frequent repaving of the ground surface. The lack of aroid pit cultivation and rocky nature of the landscape indicates that the islet was probably occupied by small groups for short periods of time, perhaps as a staging area in which to exploit the outer reefs and troll for pelagic species. A basal radiocarbon date suggests late prehistoric occupation. DATING The selection of samples for radiocarbon dating was constrained by the limited number of excavation units. However, the objectives for securing chronometric dates were to: (1) obtain a set of dates for the basal cultural layer which would establish initial occupation in the location of the units, but not necessarily of the site itself; (2) determine the relatively oldest portion of the site along a transect that would suggest the direction of site expansion over time and how this may correlate to islet formation; (3) contrast the basal habitation dates between the largest islet and at least one site on a small islet; and (4) date a sample of beachrock in reference to an elevation transect for establishing a chronometric benchmark for this topographic feature that may have implications for human colonization of the islet. Seven radiocarbon samples were selected from clear stratigraphic contexts: four from the largest habitation on Ujae Atoll, site 1; two from the small habitation site on Bock islet; and a beachrock sample from the ocean side transect on Ujae islet. All samples were proc- Site 3 Site 3 Unit 1, W profile Unit 1, E profile 2ress BOSS - pr ‘obable base of historic dense gravel dense oven layer i) dense gravel = co) sterile sand layer dense gravel dense oven stones eS) @) e 9 Beta-76019 200 + 60 BP (wv) DSS. medium-coarse sand beach sand and cobbles e charcoal FIGURE 13. Profiles of unit 1, site 3, Bock islet. essed by Beta Analytic Inc. Pretreatment of the carbonized material consisted of mechani- cal cleaning with gentle crushing, followed by washings in de-ionized water and removal of any rootlets. Acid washes were used to remove carbonates—a potential problem with sam- ples recovered from calcareous sands on atolls—and alkali washes removed secondary organic acids. The beachrock sample was processed in a similar manner. All procedures went normally. Table | presents the details for each dated radiocarbon sample. When possible, discrete combustion features were selected for dating of known material such as coconut shell or husk and Pandanus keys (or drupes). Unidentified wood charcoal was part of two samples and the sole material for one sample. Ideally, it is best to get wood charcoal identified prior to dating to eliminate the possibility of “old” wood that can produce dates not related to the target event (Dean 1978). A thumb-sized piece of wood charcoal was submitted from an oven (Beta-79582) to Gail Murakami of International Archaeological Research Institute, Inc. for identification. The sample could not be identified due to a lack of adequate refer- ence material. It is prudent that atoll reference plants be collected to facilitate wood char- coal identifications which would not only furnish better control for radiocarbon dating, but provide insights into plant use and, perhaps, environmental change during prehistory (Hastorf and Popper 1988; Weisler and Murakami 1991). The oldest date for the site 1 transect was at unit 2 (AD 256-542), just lagoonward of the Cyrtosperma pit concentration. This area, and similar settings on other atolls, has been shown to produce the earliest habitation dates. It is likely that upon discovering Ujae, the colonists explored the atoll and located the largest islet. It is near the center of Ujae islet that the Ghyben-Herzberg lens would be the largest and thickest and the prime locale for 19, ‘aw Bugunod papua}xa uantb sajdwes (q) *(€66T) sawlay pue JaAinjs Jaye payerque (e) "}RODJeYD pOOM payuaptun = 4 ‘fay SnuDpUdg = € ‘}JaYS NUODOD = Z /ysNYy ynuUOI0D = T ‘anbiuyda} Hurep IuZewolpes Aq pazAjeue sajdwes jy O€l dV - STE DA aptsuess0 0°965 = JIOAYIPOq OL + 0752 Eats 09 + 002 w10j3e)}d yaa1 aAoge wT “sy aely 18856 SG6L - LEST OV uaro — (q) 9" y 09 + 002 €°82- 09 + 092 (squd ¥21-2E1) III/t-€ 6109Z SS6L - Y69T dV UsAO €°29 T 0S + 0€ Vales 0S + 04 (squwi 4/-29) III/T-€ €8S6Z uJapow UdAO 99 ¥'€ %oL°00 + €° LOT 8°92- %L°00 + O'TOT (sqw 0€-02) I/L-1 C8562 uJepouw pasuadsip Ale €'2 %9°00 + 9°00T 8°S¢e- = %9°00 + 7 00T (squid 04-0€) III/9-T T8562 6¥VL - 68¢eL AV UBAO (q)¥"€2 % 'T 0+ 095 0°8e- 01+ 019 (squid ¢/-S9) AI/S-T 8L09Z 24S - 992 AV nyts ul Cube T 09 + 099T c.9¢- 09 + 089T (sqwo 611) II/2-1 SY8yL (e)abuey 0 2 }x9]U0) 1ybiam jeuajew jeuoluaAuo) Do/ Der Dr BIUBLUBADIg “ON BJAG poezeiquey “SuotjeULWajap abe uogieroipel yjozy ae[f “LT a]GeL 20 aroid pit cultivation. North of unit 2, units 5 (AD 1289-1449) and 6 (modern) produced progressively later dates documenting a clear chronological trend from near the center of the islet to the lagoon. While later habitation dates may reflect differential site use and population expansion, islets do build lagoonward (Wiens 1959) and over a period of several hundred years, new land was undoubtedly forming thus providing additional space for set- tlement. In the context of islet formation, dating the ocean side beachrock formation establishes when this topographic feature existed (BC 315- AD 130). Admittedly, the sample is not reckoned precisely to modern sea level, but the date may establish a terminus a quo, or starting point, in which to determine when the islet could have supported human colonists. It is unlikely that people colonized Ujae prior to the beginning of the first millennium AD. This is in agreement with the earliest dates for Ebon, Kwajalein (Shun and Athens 1990), Majuro (Riley 1987), Maloelap, and Utrok atolls. Few, if any, archaeologists accept the 3000+ BP dates reported by Streck (1990) for Bikini Atoll, where he has yet to provide the cultural context and complete stratigraphic details of the samples. If the earliest habitation dates are on the largest islet of each atoll, the smaller islets should have later dates of use. While I am not suggesting that people did not use all the islets from initial colonization, the routine, and more permanent use of these locales, was probably after a period of settlement build-up on Ujae islet. As seen today, residents of the main village venture farther afield for bottom fishing in the lagoon, only after stocks are reduced close by. Diminutive islets, especially without a fresh water lens, could only be used for short periods of time and, consequently, may have only supported small groups while capturing birds, collecting shellfish, and fishing the adjacent shores. In summary, Ujae Atoll may have been sufficiently above sea level about 2000 BP and, having a large enough land mass and resulting Ghyben-Herzberg fresh water lens, could only then support a permanent human population. The earliest habitation date of AD 256- 542 establishes that people settled sometime after the first few centuries AD and used the smaller islets more intensively in later prehistory. The earliest habitation date does not document initial colonization, but suggests that the earliest cultural deposits on Ujae have not been located. ARTIFACTS It is on low coral atolls that shell resources are the primary raw material for tool manu- facture. With an absence of local volcanic or continental rocks, the fashioning of shell into a wide range of functional and ornamental artifacts was taken to extreme. For example, in the Marshall Islands, adzes were made from at least seven shell taxa: Tridacna maxima, T. gigas, Cypraecassis rufa, Cassis cornuta, Lambis lambis, Conus sp., and Terebra maculata. Vegetable peelers and scrapers—modern examples still in use today—were made most frequently from the gastropod Cypraea tigris and probably C. mauritiana and C. maculifera, as well as bivalves including Anadara sp., Asaphis sp., and Pinctada margaritifera. A shell weaving tool was fashioned from Tridacna. The large, flat pearlshell bivalve, Pinctada margaritifera, provided ample raw material for making a range of single-piece fishhooks and trolling lures; Turbo was also used, but to a much lesser extent. The widest spiral near 21 the apex of large cone shells (Conus leopardus and C. literatus) was shaped into rings that were displayed on arms and hung within greatly distended ear lobes. The raw material for other ornaments, including beads and pendants, was supplied by red-colored Spondylus valves, Conus, and possibly Strombus luhuanus. Abrading tools were made from coral and pumice. The archaeological assemblage of Ujae artifacts included 35 tools and ornaments of indigenous manufacture (excluding manuports). Most of the specimens are relatively large surface artifacts collected during the survey; however, a few artifacts were subsurface finds (Table 2). I describe the collections under broad functional classes. Shell Adzes Adzes made from the shells of Tridacnidae—the largest bivalve known—have a wide distribution in the western and northern Pacific and are the most common shaped artifacts found in Marshall Islands habitation sites. The 23 specimens reported here are surface finds Table 2. Artifacts from sites 1 and 3, Ujae Atoll, Marshall Islands. Type/Class Site 1 Site 3 Total Historic Button 2 0 2 Ceramic 3 0 3 Glass 68 2 70 Metal 154 2 156 Nail 17 0 17 Plastic 4 0 4 Subtotal 248 4 252 Indigenous Abraders Coral 1 0 1 Pumice 5 0 5 Adzes Cypraecassis 7 0 7 Tndacna gigas 0 4 Tndacna maxima al 0 11 Trdacna sp. 0 Needle Bone 1 0 al Ornaments Conus ring 3 0 3 Tectus ring 1 0 1 Manuports Non-oceanic rock 3 0 3 Oceanic rock 1 0 1 Pumice 187 1 188 Worked Shell 0 Pinctada sp. 1 0 al Subtotal 226 1 227 Total 722 9 731 22. from site | on Ujae islet. Because of the relatively small sample size, I followed the classi- fication system devised by Kirch and Yen (1982:208-232) for their analysis of 234 Tikopian shell adzes. Three taxa were identified in the Ujae collection: Tridacna maxima, T. gigas, and Cyraecassis rufa, and the adzes are described under each taxon. Tridacna maxima is the most commonly-used adze material found in the Marshall Is- lands and while present throughout the entire sequence, it is more often associated with late prehistory. Fifty percent of the Ujae adzes identified to species are T. maxima (Fig. 14a). Adzes were commonly shaped from the valve’s dorsal region where the long axis of the adze blade was perpendicular to the radial flutes (see Kirch and Yen 1982:fig.84). This is the region of the valve where the longest adze could be obtained. Also within the dorsal region, 9 of 11 (82%) of the adzes incorporated the area demarcated by the retractor muscle scars undoubtedly because this surface is more parallel to the valve’s exterior than any other portion of the shell. Consequently, less grinding would be necessary to derive the final adze shape. It may also be that the shell material is denser and more durable at the retractor muscle scars and, consequently, may have produced stronger adze blades. Of the three adzes where valve side could be determined, all blades were made from the ventral valve (Fig. 14b and c) and measured up to 88.63 mm long. All 7. maxima adzes that could be assigned a type were type 3 (Kirch and Yen 1982) which is a more specific classification of Rosendahl’s (1987) TRI-EXT (exterior region of the Tridacna valve, literally, Tridacna exterior). The Ujae adzes typically have 50% or more grinding on all surfaces, are quadran- gular in cross-section with rounded or blunt-shaped butts, and most frequently have slightly curved cutting edges (in plan) that are flat (see Tables 3 and 4). Four T. gigas adzes (and possibly a fifth, Fig. 15a) display the widest range of variabil- ity in size and shape. This is predictable since T. gigas reach more than 1.0 m in posterior- anterior length and 10 cm thick, supplying a virtual “blank slate” for adze design. From comparisons with modern reference specimens of 7. maxima and T. gigas, it is readily apparent based on adze size (Fig. 15c) and thickness (Fig. 15b) that these four adzes can only have been fashioned from T. gigas. The width of SA 17 (Fig. 15c) exceeds that of average T. gigas flutes and was therefore made from the hinge portion of the valve. SA 2 (Fig. 15b) could have been made from the hinge or flute region of thick individual valves. As a class, these adzes typically exhibit 100% grinding and are the longest in the assem- blage at 134.22 mm. Cross-sections are plano-convex and quadrangular with blunt or pointed butts, slightly curved or straight cutting edges with concave or flat bevels; these adzes are similar to Kirch and Yen’s (1982) type 6 and 8 made from 7: maxima. T. gigas adzes are generally the oldest forms of shell tools in culture-historical sequences from the western and northern Pacific. It is interesting to note that SA 17 (Fig. 15c) may have a residue on the proximal region of the back that may be the result of hafting. The second most numerous group of Ujae adzes were manufactured from the lip or whorl of the Bullmouth Helmet shell Cyraecassis rufa. A number of authors have confused this taxon with the Horned Helmet shell (Cassis cornuta) when referring to some adzes made from the lip or whorl (e.g., Beardsley 1994:photo 15; Kirch and Yen 1982:fig.91g; Shun and Athens 1990:fig.3 specimen 2). Although adzes, chisels, and gouges were made from the lip of Cassis cornuta and Cypraecassis rufa, only whorls of the latter taxon where fashioned into adzes. I have seen no adzes made from the whorl of Cassis cornuta either in 23 FIGURE 14. Tridacna maxima shell adzes. All specimens made from the dorsal region of the valve, perpendicular to the radial flutes. a and b = finished forms; c = adze preform flaked to rough shape, ready for grinding. Artifacts b and c made from the ventral valve. “WO} Ul JE}LWIs aq 0} Aeadde sazpe sobib +) aelf ay} ybnoyye ‘owixow DuIDpUY 0} 1aJa1 g pue 9 sadAy (Z2g61) UaA pue YDILy ‘apis = s “y2eq = q ‘JUOIJ = 4 :sU0L}eD07 Bulpuud ‘jeIQUAaA = A ‘uoLHal jesuop = ip :uoLBay yJays “suoeLAaiqge , OL 3e)4 pean 0S-0 s ‘q poum - X@AU0d-aAeIU0D = abpa+piwi peysiuy 92 VS _ pfns sissodan1dA 6 aaequod paund 0S-0 s‘q’s dy pajutod xeAuod-our}d 3]0yM paysiuy 42S {ni sissodap1dA) OL ey pavind 0S-0 s‘q yuoum — X@AU0I-aAeIU0D «= abpa+piw paysiuy 02 VS -afnd sissodan1dA) OL 34 panino Ay,ybuys 0S-0 s‘q youm paquiod xaAuod-aAedu0D 2]04M peysiuy 6L VS of sissodanidA OL ey paaind Ajjybus 0S-0 s ‘q youm pajuiod xaAuod-aAeu0d a]oum paysiuy GL WS dfn sisspdapidA) 6 ef panund Oot s‘q’5 dy punos xaAuod-oue}d 3]0ym paysiuy GS pfni sisspdan1dA) OL 3ey4 pavind 05-0 s ‘q youm _ X@AuOd-aAeIU0D == abpa+piw paysiuy € VS {nd sisspdapidA) é8 = = Oot s‘q’ é papunos é yng so paysiuy = T2 S sobi6 bu2vpuy 8 3ey4 yybtes4s Oot s ‘q ‘5 é paqutod 4e)nbueupenb 3]}04M paysiuy LL YS sobib pudppuy 9 _— ~ OOT segue é papunoi xaAuod-our}d plw+}3nq paysiuy OL vs spbth pudppuy 9 aaequod padind Ayybuys OO SG 4apen é qunjq xaAuod-oue)d 3)oym paystuy 2 VS sobtb pudppuy 8 3e)4 pavino OOL SeCiay é _ 4eynbueipenb abpa+piw peysiuy SL VS ‘ds pudppuy € _— _— OOT-0S Seq 4p papunoi Je)nbueipenb piw+}3nq paystuy G2 VS DwIxXbu DUIDpUY € ey pavind Ayybus OOT-0S Sate (ie 4p yung Je)nbueipenb 3)oum paysiuy €2 VS DuILxDW DUDDpLY é€ _ _ 0 — dp ‘A pepunos Jeynbuespenb 3}0uM Wojaid 9L VS DWIXDW DUIDpU € ey pavund Ajybus OOT-0S s‘q‘s dp ‘A papunos Je)nbueipenb a]oum paystuy, 4L YS DWIXDW DUDDpLI| _— 34 qybres3s OOI-0S s‘q ‘5 4p — Je)nbueibenb abpa+piw peystuy €L VS DWIxXDW DUIDpU — yey pavund Ajybus OOT-0S tap a dp ‘A _— 4e)nbuespenb abpa+piw paystuy 21 VS = DWLXD DUIDpU € -_ — OOT-0S s‘q’j up qunyq 4Jeynbuespenb piw+39nq paystuy IL VS DWixbw DUdDpL) é€ 3e)5 pana OOT-0S s‘q’‘J ip - 4eynbueipenb abpa+piw paysiuy 6 VS DWIxDW DUDDpL € ey WyBbies3s OOT-0S 5 Hap al 4p papunos Jeynbueipenb a]oum paysiuy LVS Dwixpw pudDpuy € — - OOT-0S s ‘q's 4p papunoi Je)nbuespenb piw+39nqg paysiuy 4 WS _ DUILXDLU DUDDPLY — — — OOT-0S Suge 4p _ Jejnbueipenb ~—uolqdaspiw peysiuy LVS DWIxDW DUIDpLY uoeoytsse)) ]eAeg UP}g UT % Burpuuy uolze07 uoibay adeys uoyzas uolpyog abejs “ON uoxey (2861) UaAR Yu ebp3 buyin) Butpuug = 84S yng “$801 bojeye) 24 ",9azpe yjays aelf jo saynquyye ajasosiq “€ a}qGe) 25 "YIpiM yJod/yypiM JuLodpiw jo o1Ves = Jade) ‘ybuaj/sseuy1y} JuLodpiw jo o1je1 = 7) ouey ‘yybuay/yypim uLodpiw jo oe = 7M Oey ‘ssaux2143 JULodpiw/yIpM JULodpiLU JO OlJe1 = 1M OL]eY “suoteaiqqe , = = = = 0” = = 81°S = = = = €€°2 = = 96°S — = €9°S 68°S GOST = 2G = = L6°2 12°0 82°0 ¥0°L SY 8°99 €£°OL GS°22 L8°L 6£°E2 €S°02 22°€8 ey'T 910 25°0 LEE 09 169 0€°Z 6LIL OL'82 = BL'6E S8°SY 68°S/. = = = = a = = — 9915 — = = = = a 9ELL Gy = = Gore = Oy'ly 00°%y = Gee 60°0 9€°0 ly GS L'ze 718 22'L EL°6 59°62 OE "LE S218 = “= = €9°E Ov = = 18°9 = L6"¥2 EL°G2 = 6L°2 STO Ly°0 bce GY 6°S62 88°ST 22°02 9q°22 = 8°29 80°2L 22'VEL bike a) £40 Or'y = EYL 66°9 Teer [eee aeS 18°6€ 81921 yLE Oa) 27°0 Loy Ov L6E GEL Els £8°8 90°€€ 80'ly 18°8Z 69'T 200 L¥°0 60°Y 0s VEL 1E°6 L201 92°82 LO'LY 25°67 €9°88 = = = 9L°9 0S = = 90°S = O2"vE OLE = = = = 82°L Gy = = 40°9 = 86°EY 20°07 = 16'T = = 25'S = = 40°9 20°9 BELL —s L2"€€ = _ = = = = == = 65°82 = 80166 == = = _ om = 69'y 0€ = = yUL = 64°E€ EVE = SEZ 9L°0 770 OL'2 Ov 9°SY 72'9 96°11 O/;ET =smecace EL'EE 9S°€L ole St0 92°0 ELT 0” 212 LES S0°OL €y°9 SELL OL"61 94°99 66'T = = 9€°S = = 16°L 1y'8 e922 = ¥0"GY oe = = = = 0S°9 Ov = == 0S"y = S262 an = ey 1E0 62°0 76'0 0S 1°89 8L°81 G2"y2 80°91 68°22 8°12 9L°8L = a = Ivl = — = 08"y = 85°GE = = jade) Th 1™ IM — ajbuy a6pz yybiay yng qurodpiy Ned qutodpiw abpy bun) §=—-y3uay onjey oney oey Buiyan) SSAUxILY] —- SSBUADLY| YPM 4IPLM 4IPLM 92 WS G2 VS 92 VS €e WS T2 VS 02 YS 61 WS 81 VS LT VS 91 VS ST WS YL WS €L VS el WS IT WS OL VS 6 VS LYS S VS yYS € WS e WS TVS “ON bojeje) *,S9Zpe pays aelf jo saynquzje snonuquo) *y age 26 FIGURE 15. Shell adzes of Tridacna gigas typically display a wide range of cross-section form. 2 publications or in the Alele Museum (National Museum of the Marshall Islands) collec- tions. Inspection of whole and fragmentary reference shells shows clearly that the whorls of Cassis cornuta are much thinner with an uneven surface. Five of seven (71%) specimens were made from the whorl of Cyraecassis rufa shells at least 138 mm long (Fig. 16c and d). Grinding is minimal on the back and sides and is not required on the front surface. All specimens are concave-convex in cross-section with curved to slightly curved, flat cutting edges at an angle of 40-55° (see Tables 2 and 3). These specimens are type 10 of Kirch and Yen (1982). Two adzes made from the lip region are both plano-convex in cross-section and are classified as type 9. Grinding is minimal (Fig. 16a) to 100% (Fig. 16b) The cutting edges are curved and flat or concave. Abraders These artifacts were used for shaping and smoothing relatively soft material—such as thin shells (e.g., Conus sp. and Pinctada sp.), wood, and bone—into tools and ornaments. A single Acropora sp. branch coral abrader has a rounded and smooth surface that would be ideal for shaping the interior bend of pearlshell fishhooks. It measures 71.22 mm long and FIGURE 16. Adzes made from the lip region (a and b) and body whorl (c and d) of Cypraecassis rufa. 28 weighs 6.0 gms (Fig. 17f). According to Rosendahl, who made extensive surface collec- tions throughout the Marshalls, and my examination of the artifact collections in the Alele Museum, Acropora sp. branch coral abraders are much less common than the Porites sp. forms that are angular and blocky (cf. Rosendahl 1987:fig.1.67, a, c, and h with Fig. 17f). Four pumice abraders (mean length = 40.92 mm; mean weight = 4.28 gms) each have multiple worn facets that were used to shape flat and convex surfaces (Fig. 17g, h, and i). Fosberg (unpublished, cited in Sachet 1955:12) reported that a large black, coarse-grained, hard pumice was seen in use as a whetstone for machetes on Bock islet, Ujae and a similar piece was found inland in the northeast region of Ujae islet. It is plano-convex with use- wear on the flat surface and measures 106.20 mm long and weighs 158.6 gms. Worked Pearlshell Only one piece of clearly worked shell was recovered. A 28.10 mm long specimen (16.08 wide and 4.72 mm thick) of Pinctada sp. may be the by-product of fishhook manu- facture (Fig. 17c). It is slightly ground on a portion of the interior surface. Ornaments Shell rings are the most ubiquitous ornaments found in the Marshall Islands and have been referred to as arm bands, bangles, or circular units, to name just a few (cf. Kirch 1988). The rings were variously worn on lower and upper arms as well as within greatly stretched ear lobs. The rings were most commonly made from the largest cone shell (C. leopardus or C. litteratus), although Tectus and probably Trochus were also used. Portions of four rings were recovered from Ujae, displaying three styles made from Conus and Tectus. The largest ring fragments—with projected diameters of 60 and 80 mm—display up to four neatly-incised parallel grooves (Fig. 17a) and are similar, but not identical, to a ring re- ported from Ebon Atoll (Rosendahl 1987:fig.1.761). The illustrated Ujae ring measures 16.69 mm wide and 3.00 mm thick, while the second one measures 5.29 mm thick with an incomplete width. Another Conus ring has raised parallel margins and an estimated whole diameter of 50 mm, a width of 13.91 mm, and 2.26 mm in thickness in the center (Fig. 17b). Despite being only 18.41 mm long, a shell ring fragment of 7ectus has a reconstructed diameter of 60 mm; it has no design. A small Terebra cf. dimidiata shell (length = 83.18 mm) has a hole chipped near the aperture to facilitate suspension as a pendant (Fig. 17e). No other Zerebra shell pendants have been reported from archaeological contexts in the Marshalls. Bone Tool Possibly made from a mammal long bone, this elongate, flat, pointed artifact measures 76.57 mm in length and, at the center, is 7.72 mm wide and 2.65 mm thick; it weighs 1.6 gms. The proximal portion is biconically drilled, while the pointed end is worn and blunted probably from use as a thatching tool or needle (Fig. 17d). It is covered with scratches oblique to its long axis and has a polished appearance overall. This is the only known example of this artifact type from the Marshalls. 29 FIGURE 17. Assorted artifacts from Ujae excavations a = Conus shell ring fragment with four neatly incised grooves; b = Conus shell ring fragment with slightly raised edges; c = worked pearl shell (Pinctada margarttifera); d = possible thatching needle made from mammal (?) bone; e = Terebra pendant; f = Acropora sp. file or abrader; g - i = pumice abraders with flat and rounded used surfaces. 30 Manuports These artifacts include physically-unmodified items that are foreign to their place of cultural deposition. Included here are pieces of pumice that were probably collected along the lagoon edge or ocean shore and brought to the habitation site for shaping into abraders or simply crushed and added to garden plots to enrich the soil with iron and trace elements (Sachet 1955). Because pumice was found only in the cultural deposits, and not the sterile subsoil, it is assumed here that the material was humanly transported; it can, however, occur naturally in the interior of smaller islands. Some 187 unshaped pieces weighing 74.6 gms were recovered from site 1, while site 3 contained only one specimen weighing 0.2 gms; all material was from subsurface contexts and has identical fine texture and absence of large inclusions; Munsell color of unweathered interior surfaces was 2.5Y 7/3 (pale yellow). A piece of pumice collected by Fosberg on Ujae in 1952 had an iron content of 2.7% (Sachet 1955:21). A recent geochemical analysis of Ujae pumice is presented in Table 5. Four additional manuports were found as surface artifacts and, unfortunately, cannot be placed chronologically. They are, however, important for documenting the availability of non-local rocks that, on other atolls, have been used for abrading tools, hammerstones, and oven stones. None of the Ujae non-pumice manuports exhibit evidence of physical modifi- cation. There are two specimens of metamorphosed quartz sandstone (65.7 and 238.6 gms), an altered rhyolite lava (261.5 gms), and a highly weathered volcanic rock (117.4 gms). X- ray fluorescence analysis and thin-section petrography were conducted to gain some in- sight as to the geological origins of these rocks (Tables 5 and 6). The sandstone and rhyolite specimens come from mainly mature island arcs or continental margins such as Japan, western United States, and New Guinea (John Sinton, personal communication, 1994). The weathered volcanic rock originated from an oceanic island, the nearest is Kosrae, some 540 km (290 nautical miles) southwest. Other oceanic rocks have been found in the Marshalls (e.g., Namu Atoll, Mason 1947:10; Spennemann and Ambrose 1997), yet the Ujae speci- men is quite weathered and soft and not ideal for tool manufacture; there is no evidence suggesting use as an oven stone. It is unlikely these manuports were humanly transported from their place of geologic origin, as they may have arrived in the roots of drift logs. It is also possible that the metamorphosed quartz sandstone is discarded ship ballast from the German era (1857-1914) when vessels plied the Marshalls to acquire copra. FAUNAL MATERIAL Fish Understanding prehistoric fishing strategies in the tropical Pacific and how practices may have evolved over time has been a subject of major inquiry for the past few decades. There have been numerous detailed studies of portable technology such as fishhooks (Emory et al. 1959; Sinoto 1959, 1962; Weisler and Walter 1999) as well as summaries of the impressive fish ponds and traps of Hawai‘i (Kikuchi 1976; Summers 1964). How and why fishing strategies changed over time has been the subject of numerous papers from New Zealand (Anderson 1997; Leach and Anderson 1979; Leach and Boocock 1993), the Cook Islands (Allen 1992; Walter 1998), Hawai‘i (Kirch 1979), the Marquesas (Rolett 1989; Leach et al. 1997); the Societies (Leach et al. 1984), a few areas of Micronesia (e.g., Leach et al. 1996), and Near Oceania or the western Pacific (Butler 1994). Ethnographic studies 31 Table 5. X-ray fluorescence analysis of oceanic and non-oceanic rocks from Ujae site 1. Artifact Number S.A. 6 S.A. 8 S.A. 27 S.A. 28 Pumice Normalized Results (wt%) Si0, OTT, 75.57 45.02 69.27 67.27 Al,03 12.99 13.30 14.28 16.09 15.43 Ti0, 0.766 0.163 +6.04 0.888 0.390 FeQ* 4.88 1.44 11.53 5.47 3.88 MnO 0.089 0.016 0.207 0.048 0.119 CaO 1.62 0.33 12.02 0.63 5S i. MgO 2.33 0.09 5.58 2.06 1.17 K,0 2.05 5.06 2.85 3.54 1.09 Na,O 3.36 3.97 1.48 1.84 4.05 P20. 0.153 0.068 +1.00 0.17 +1.129 Trace Elements (ppm) Ni 24 10 100 37 5 Cr 99 2 19 69 5 Sc 17 2 28 22 18 V 72 0 397 84 39 Ba 233 611 849 492 184 Rb 82 174 68 130 14 Sr 220 100 762 106 597 Zr 286 103 450 193 83 \f 27 38 38 33 19 Nb 12.6 15.9 +89 22.0 1.4 Ga 9 10 21 17 13 Cu 7 1 79 5 5 Zn 63 12 +130 42 70 Pb 10 ilal 0 6 3 La 14 52 53 18 0 Ce 66 101 119 72 15 Th 8 24 4 12 0 Major elements are normalized on a volatile-free basis. *Total Fe is expressed as FeO. “ou +" denotes values >120% of the highest standard. For analytical procedures see Hooper et al. (1993). (Dye 1983; Kirch and Dye 1979) have been especially useful in expanding knowledge gained solely from the analysis of artifacts, fish bones, and their spatial relationships. Of importance here is the excellent exposition on Belauan traditional marine lore by Johannes (1981). And I personally cannot argue strongly enough for participant observation which has been vital to my understanding of prehistoric fishing. For example, some would argue that parrotfish (Scaridae), with its robust mouth parts, is an important family when consid- ering its contribution to prehistoric subsistence. However, having spent more than two years on more than half a dozen atolls in the Marshalls, it is obvious that this family is over- represented in the archaeological assemblage because its hard mouth parts—used for iden- tification—preserve well and bear little resemblance to the rank-order abundance by family of what is caught today. 32 Table 6. Petrographic descriptions of oceanic and non-oceanic rocks from Ujae site 1. Artifact No. Description S.A.6 Metamorphosed quartz sandstone, now a slightly foliated, partially recrystallized, quartzo- feldspathic schist. There are lots of strung-out quartz and feldspar clasts about 1- 1.5mm long in a foliated matrix of either biotite or possibly stilpnomelane, quartz, feldspar and some zircon and apatite. This is a low grade (greenschist, about 350°C) metamorphic rock of an originally quartz-rich sediment. Such rocks come from mature arc or continental settings. S.A.8 This is an altered rhyolite lava with lovely quartz and altered feldspar phenocrysts about 2 mm in size in a largely altered, very fine-grained groundmass. The feldspars are completely replaced by a cloudy, cryptocrystalline mica, probably mainly sericite. There are also some iron-rich olivines in this rock. Rocks like this come from arcs, mainly mature island arcs or continental margins, e.g. Japan, western US, New Guinea etc. S.A.27 This rock also is a metamorphosed quartz-rich sediment but it has quite a different texture than S.A.6. There are subrounded clasts of quartz and feldspar about 0.3-0.5 mm in a matrix of pale brown biotite, a little muscovite, apatite, quartz and feldspar. The rock is not strongly foliated or schistose. The temperature of metamorphism is similar as for S.A.6, but the texture is quite different (different stress regime). Also from some mature arc or continent. S.A.28 A volcanic rock without phenocrysts, there are microphenocrysts (~0.3 mm) of Ti-rich clinopyroxene in a vesicular groundmass of Ti-clinopyroxene, plagioclase and magnetite. Rocks like this come from oceanic islands. I would like to illustrate the importance of local fishing knowledge when interpreting a fish faunal assemblage. On Utrok Atoll, from late November to early February (1996-97), goatfish (Mullidae) and rabbitfish (Siganidae) made up more than half of the fish consumed in the village. It was quite common to capture schools numbering in the hundreds by seine nets positioned out from the shoreline. In fact, I ate Yellowstripe goatfish (Mulloides flavolineatus and M. vanicolencis) nearly every day and it wasn’t even the peak season for its occurrence (Myers 1991:148). Both mullids and siganids have relatively fragile mouth parts and their limited occurrence in archaeological deposits is probably due more to taphonomic processes, than prehistoric subsistence practices. On Airik islet, Maloelap Atoll, I witnessed in November (1993), a fishing technique that has virtually no archaeological visibility—both of the tools used for capture and the result- ing fish bones that may become incorporated into cultural deposits. In the very early morn- ing hours, when the water is calm, about a half dozen men position themselves along the lagoon shore, each spaced about 3-5 m apart, standing on the wet sand. Each person holds at opposite ends, a tightly braided, 2 m coconut leaf that is bent in a U-shape. Constantly watching the inshore waters, the men scan for predatory carangids that are chasing schools of juvenile mullids, herding them ever closer to the dry shoreline. Once the school is forced into water less than 20 cm deep, the young fish flee the predator by breaking the surface in unison thus alerting the men to their presence. Each man then swoops down with his braided 33 coconut leaf forming a barrier around the school, open to the wet sand. The leaf is then dragged along the bottom, pulling slowly towards dry sand. A couple kilograms of fish can be trapped by each fisherman. The fish are then consumed whole. There is, of course, no visibility of this technique in the archaeological deposits. The last example I will mention is that recorded on Ebon Atoll in 1995. The Marshallese word, ajilowdd refers to a “herd of bonitos that enters [the] lagoon and can’t find its way out” (Abo et al. 1976:307). Etna Peter, longtime Ebon resident, described, by word and picture, how large schools of skipjack tuna (Katsuwonus pelamis) would enter the lagoon. Each family of the village then gathered coconut leaves which were woven to form a loose braid several meters long. The weave was loose enough that leaflets extended in all direc- tions around the braid. By securing numerous sections of coconut leaf braids (one produced by each family), the length was extended to a couple hundred meters. The “rope” was then loaded onto several canoes with the goal of getting the school positioned between the la- goon shore (at the main village) and a half-circle line of canoes. The canoes moved slowly towards the shore, scaring the tuna to shallow water where they were speared by waiting villagers. As I have witnessed on numerous occasions, fish are hesitant to swim under thick ropes lying on the surface. Although this technique would not be visible archaeologically, the presence of scombrid bones in archaeological sites from the Marshall Islands, is not an automatic proxy for inferring the trolling fishing technique. The archaeological fish bones from Ujae Atoll were identified to the lowest taxonomic level possible using an extensive reference collection derived in large part from the Marshall Islands and excellent illustrations of scarids in Bellwood (1994). For those families includ- ing Carangidae, Labridae, and Scaridae, sufficient reference material permits identification to genus. This is especially important since carangids inhabit a wide range of environments and can be captured by varied techniques. Scarid and labrid bones identified at the genus level can sometimes reveal their habitat preferences which may relate more precisely to the ecology of fishing strategies (Bellwood 1994; Bellwood and Choat 1990; Myers 1991; Randall 1996). Some 4,473 fish bones were identified, 575 (12.7%) to family or below (Table 7). The five-paired mouth parts (maxillary, premaxillary, dentary, articular, and quadrate) along with other post-cranial bones were used for identification. These latter elements included dermal spines of diodontids, scutes and pterygiophores of carangids, scales of balistids, and caudal peduncle of scombrids. The assemblage was also separated into cranial bones, spines, and vertebrae to gain a better understanding of the assemblage structure. For example, there are more than twice as many vertebrae at site | than there are at site 3, and this difference may relate to variable processing techniques, consumption patterns or, perhaps, taphonomy. While I make no pretence that the fish bone assemblage is a random sample, it should broadly represent what is present in the archaeological deposits. Some 19 families were identified which is only about five families less than much larger samples from Ebon, Maloelap, and Utrok atolls. Since there was no fishing gear recovered from the excavations, it may be that seine net fishing from the ocean side, but mainly from the more protected lagoon shoreline, was a dominant strategy. With the exception of Belonidae, Cirrhitidae, Diodontidae, Lethrinidae, marine eel, and Scombridae, the majority of fish could be captured by net; this is 95% of all 34 Table 7. Number of identified specimens (NISP) for fish bone from sites 1 and 3, Ujae Atoll. Taxon 1 33 Total Acanthurnidae 10 3 13 Balistidae 20 20 Belonidae il 1 2 Bothidae 1 2 3 Carangidae Caranx 2 1 Elagatis Al Selar 1 1 Total Carangidae (all taxa) 31 20 51 Cirrhitidae 1 1 Diodontidae 14 14 Elasmobranchii al 1 Exocoetidae 1 1 Holocentndae 6 6 Labridae Labridae (to family only) 2 4 Bodianus 2 Cons 1 Thalassoma 2 il Total Labridae (all taxa) 7 5 12 Lethninidae Lethnnus 1 Monotaxis 2 2 Total Lethrinidae (all taxa) 3 3 Marine eel 1 1 2 Mullidae 2 2 4 Nemipteridae 1 al Ostraciotidae 1 15 16 Scandae Scandae (to family only) 52 34 86 Calotomus 1 2 3 Scarinae 53 13 66 Scarinae (not Scarus) 6 4 10 Cetoscarus bicolor 9 2 ilal Hipposcarus longiceps 17 16 33 Scarus/Hipposcarus 91 27 118 Scarus spp. 22 11 33 Scarus gibbus 4 4 8 Scarus oviceps 2 2 Total Scaridae (all taxa) 257 113 370 Scombnidae 2 2 Serranidae 48 5 53 Total identified 404 aya 575 Total unidentified 2672 1226 3898 Total cranial 1762 919 2681 Total spines 279 247 526 Total vertebrae 1035 231 1266 Total bones 3076 1397 4473 % identified 13.1 12.2 127 35 identified bones as quantified by number of identified specimens (NISP). It is often the case that one family can be captured by a number of techniques. For example, just about any fish can be speared under the right circumstances. Balistids, cirrhitids, lethrinids, and serranids are often caught by hooks dropped from a canoe off the reef slope. While bearing in mind taphonomic processes and the relatively small sample size—only Balistidae, Carangidae, Scaridae, and Serranidae are represented by more than 20 NISP each—there is every rea- son to suspect that the prehistoric residents of Ujae used a varied set of capture techniques as practiced by individuals, small groups, and, at times, by most of the village in a co- operative fashion (such as tuna fishing described above). Certain taxa do stand out as repre- senting a certain capture technique more than others. Belonidae are almost exclusively taken by trolling as well as the carangid, Elagatis, which is taken inside the lagoon, while larger individuals are caught along the ocean side reef edge. Cirrhitidae and Lethrinidae are caught most frequently by baited hook lowered from canoes on the ocean side reef or deeper lagoon. Diodontidae are usually speared. Scarids are usually netted or speared, although one time in Hawai‘i, I saw one caught by a hook baited with seaweed. Birds No other class of fauna has been more instrumental in understanding the role of humans in causing extinctions and faunal depletions on Pacific islands than birds. By some esti- mates, more than 2,000 species of land birds disappeared after human colonization of Oceania (Steadman 1995). While most attention has focussed on diachronic changes such as extinctions and the resulting description of new taxa (James and Olson 1991; Steadman 1989; Wragg and Weisler 1994), less effort has focussed on functional studies like the role of bird use in relation to the settlement landscape. What role did birds fill in the subsistence regime? How were they captured? Are there sites within the atoll settlement system that can be identified as bird procurement locales? Much has been written about prehistoric fishing through the analysis of artifacts, bones, and ethnohistoric records, but little about birds. This is understandable in terms of visibility—there are numerous artifacts associated with fishing, yet none identified exclusively for capturing birds. The Ujae bird bone assemblage, totaling 87 specimens, was analysed to understand the role of avian predation within an atoll setting. The model, proposed here, is that each wooded islet within Ujae Atoll should have supported bird colonies prior to human colonization. Indeed, on many atolls, I have observed terns nesting on rubble cays and noddies breeding on even the smallest of islets with just a few trees. It is quite likely that the largest islets with forests of Pisonia would have supported sea bird colonies in the thousands. After human colonization of the largest islet of Ujae, bird populations would have been reduced rap- idly—perhaps within a few human generations—and remnant colonies shifted to the off- shore islets. These smaller offshore islets, that could not sustain permanent human populations due to a lack of fresh water, became resource locales where small human groups visited for several days at a time for capturing birds and exploiting the adjacent marine environments. How, then, would this be visible archaeologically? The Ujae bird bones were first identified to lowest taxon by Alan Ziegler, then analyzed further by David Steadman. At the University of Otago, the bones were analyzed taphonomically by recording for each bone: species, element, portion, age, segment, length, 36 presence of burning, burning color, cut marks, gnawing, weathering, midden staining, and other modifications. Unlike most fish remains, bird bones can be incorporated into archaeo- logical sites by non-human means and it was important to remove any obvious elements that may have been incorporated into the prehistoric cultural deposits by natural deaths (Weisler and Gargett 1993). Table 8 presents the identified bird taxa from Ujae sites 1 and 3. Due to the highly fragmented condition of the elements, 64% were identified only to small, medium, or large bird. The subsurface bone was generally well preserved, while most of the elements recov- ered in the first spit or within the coral pavement of the modern village were white and clearly of recent deposition; many of these bones were of historically introduced taxa such as domestic chicken (Gallus gallus) and turkey (Meleagris galloparo), but three elements of booby (Sula sp.) were also identified. Of the total assemblage, 55 specimens (63%) were from prehistoric contexts. These bones showed no signs of burning, cut marks, or gnawing. From recent observations of Marshallese capturing, butchering, cooking, and discarding noddies (Anous spp.) and boobies, it is likely that only the proximal end of the humeri will show direct evidence of butchering. On smaller birds, such as noddies and terns, the wings are simply twisted and pulled off and show characteristic spiral fractures. While the same can be done with much greater effort for larger birds, today, wings are usually severed by cutting with a knife. The smaller, denser bones were disproportionately well-preserved and used for identifi- cation to family level. These elements included the phalanx, tarsometatarsus, carpometacarpus, tibiotarsus, coracoid, and portions of the humerus and radius. One of the more significant bone attributes was the age of the individual represented by adult or juve- nile (i.e. pre-flighted) elements. Bones of young individuals can be identified by their rough surface texture, whereas adult elements are smooth. All juvenile bones were recovered from Bock islet and strongly points to capturing young birds from their nests. Of the prehistoric bird bones identified to family, all save one, were identified to Laridae. This family includes noddies, terns, and gulls and, judging from the small size of the ar- chaeological bones, the first two are most likely. In fact, since noddies nest in the thousands throughout the Marshall Islands, it seems probable that bones of these species would be most common in the middens. The other family identified was Scolopacidae which in- cludes such shorebirds as sandpipers and curlews. Human occupation of Ujae Atoll was at least by AD 256-542 which may be several hundred years after colonization. This seems to be reflected in the total absence of extinct bird taxa which are usually a signature of the earliest occupation layers. Considering that only 1 m? was excavated on Bock islet (representing only 11% of the total subsurface sample), 55% of the prehistoric bird bones were found there. This suggests the importance of bird collecting on Bock islet and may signal a similar subsistence focus on the other offshore islets. From determining the growth stage of the individual bones, many of the birds captured on Bock were juveniles without the capability of flight. Although identifica- tion was at the family level, a small bird, such as the noddy, was the likely target prey. The low frequency of bird bones in the major habitation site of Ujae islet and high occurrence on Bock, suggests that birds may have been scarce on the main islet and birds on the off- shore islets were consumed immediately after capture; that is, they were not taken back to 37 Table 8. Summary of bone, shellfish, and other midden from sites 1 and 3. Site 1 Site 3 Total Taxon count — weight count weight count weight Bone Bird Small/medium bird _ — 5 _ 5 — Medium/large bird 36 a 7 — 43 _ Medium bird 7 — — — 7 — Large bird — — 1 — 1 a Gallus gallus* 4 _ — _ 4 — Laridae al _— 19 — 20 — Meleagns galloparo* 3 — — _ 3 — Scolopacidae — — il — 1 — Sula sp.* — — 3 — 3 — Total bird 51 _ 36 — 87 “= Fish 3076 = 1397 — 4473 — Canis familiaris* 1 Homo sapiens 27 — 0 — 27 — Lizard 2 _— 0 — 2 — Medium vertebrate 64 — 5 — 69 — Rattus exulans 8 — 0 _ 8 _ Sus scrofa* 3 Turtle (cf. Chelonia mydas) 2 — 0 -- 2 — Unidentified mammal 20 — 0 — 20 — Total bone 3305 1474 4775 Shellfish Gastropods 2722 6866.5 176 366.2 2898 UCEY 251) Bivalves 662 3857.2 196 1908.7 858 5765.9 Total shellfish 3384 10723.7 372 2274.9 3756 12998.6 Charcoal 209 84.6 458 77.9 667 162.5 Crustacea 19 6.1 5 2 24 73 Echinoderms 449 262.9 3 3.5 452 266.4 Pumice 187 74.6 it 0.2 188 74.8 * = Recovered from historic contexts. the main islet. Those smaller islets without a fresh water source were incapable of sustain- ing permanent human groups and can be seen as resource locales within the wider atoll settlement pattern. These islets, although probably visited by the first human colonists, played amore vital role in later prehistory after bird colonies were depleted near major habitation sites. Human Remains A total of 27 bones were identified as human and were only recovered from site | at units 2 and 6. As with all faunal material, Alan Ziegler analyzed the bone first, identifying the elements to nearest taxon. Mike Green, a physical anthropologist then at the University of Otago, made further observations and comments. The bones were quite fragmentary, without any signs of chemical erosion. Post-depositional breakage was probably the result of physical actions such as mixing cultural deposits through excavation for combustion and 38 other habitation features. Judging from the depositional contexts, a minimum number of five individuals are represented based on skeletal and dental fragments. From unit 2 the following were inventoried: one adolescent to small adult represented by a patella frag- ment; the left temporal (petrous) fragment of an infant; a left temporal fragment belonging to an older child or early adolescent; the permanent left maxillary tooth (M3) of someone between the ages of 17-25 years old; and from unit 6, a deciduous left mandibular (M1) tooth from a 1-2 year old child. The concentration of at least four individuals in the stratigraphic sequence of unit 2 suggests possible use of the area as a small cemetery. No grave goods could be unequivocally associated with the bones due to the mixed nature of the cultural deposits, although several shell ring fragments were found in the deposits. Other Bone Included in this category is lizard, turtle, Rattus exulans, dog (Canis familiaris), pig (Sus scrofa), medium vertebrate, and unidentified mammal (see Table 8). Only two turtle bones were recovered and this may reflect its early depletion in the first few hundred years after human colonization of the atoll. The tracks and nest sites of a few individuals noted during the survey of the smaller islets may indicate the remnants of a much larger popula- tion that has never regained its previous numbers. The situation appears to be similar to Henderson island, southeast Polynesia, where the archaeological occurrence of turtle bones is much greater in the early prehistoric period and modern sightings of the green sea turtle (Chelonia mydas) are quite limited (Brooke 1995). Bones of the Pacific rat (Rattus exulans) were found in three units from site 1. Recovery of bones of this taxon is highly correlated to the use of small mesh sieves during excavation. Because 6.4 mm sieves were used, this faunal category is clearly under-represented. One phalange of a several month old dog and two teeth and an innominate fragment from pigs four to five and nine to ten months old were recovered from historic contexts. The catego- ries of medium vertebrate and unidentified mammal may contain, amongst other taxa, a few bones of some type of smaller cetacean and those of humans, but positive identification is uncertain (Alan Ziegler, personal communication, 1995) Shellfish At least as practiced today on the outer atolls of the Marshall Islands, shellfish gathering is usually done by women and children during low tides to collect molluscs on the lagoon and ocean reef flats, and within the shallow grooves perpendicular to the ocean side reef edge during calm seas. Shellfish are collected for food as well as to incorporate into curios and certain individual shells are sold to foreign visitors (e.g., the Helmet shell, Golden cowrie, and Triton’s trumpet). The most common food taxa are Turbo, Lambis, and occa- sionally Tridacnidae. There are several species of Tridacna and the smaller taxa, such as T. maxima and T. crocea, are collected by women, while bivalves of T. gigas can weigh doz- ens of kilograms and are collected by men usually with ropes dropped from canoes in deeper parts of the lagoon. Cowries (Cypraeidae) are rarely eaten, while the Money cowrie (Cypraea moneta) is the dominant shell used in curio production. In reference to high volcanic islands, atoll shorelines—such as those found at Ujae— have a limited number of gross habitats yet a wide diversity of shell taxa are found within a 39 short distance. The nerites are generally located on the ocean side in the splash zone, Turbo are most often found along the ocean side reef edge, Lambis inhabit rubble and sand lagoon bottoms, and smaller bivalves such as Asaphis, Codakia, and Tellina prefer sandy beaches. These habitats have remained relatively stable over the past 2,000 years of human occupa- tion and changes in the diversity and richness of archaeological shellfish assemblages from the Marshalls may be related to other factors such as human predation. For example, where terrigenous runoff can prograde shorelines along high volcanic islands—burying healthy reefs that are transformed to mudflats—such radical alterations do not occur in the region. The environmental stability on Ujae Atoll permits examination of the long-term effects of humans on shellfish populations. Measures for quantifying marine shellfish are by no means standardized in archaeol- ogy. This fact is reflected in current debates (e.g., Mason et al. 1998). Some prefer identify- ing the minimum number of individuals (MNJ) by recourse to recurring attributes such as hinges on bivalves and apices on gastropods. This, however, eliminates most highly frag- mented shells which may be a significant percentage of all shellfish by weight in an assem- blage. Others prefer tabulation by weight, which may over-represent some taxa that have large individuals such as Tridacnidae. In addition to quantification issues, other factors can influence the interpretation of marine shellfish assemblages in the Marshall Islands, such as differentiating material deposited by people vs. those specimens that occur naturally in coastal sediments (see Carucci 1992 for Belau). First, there has been a long tradition of gathering coral gravel from the ocean side beaches to pave village surfaces. Along with coral, many kinds of whole and fragmented shellfish become incorporated into archaeo- logical sites and these shells can be incorrectly interpreted as food remains. Many large shellfish, including Conus, Tridacna, and Pinctada, are the raw materials for ornament and tool production. Fragments of these taxa may represent industrial waste and not food shell. Especially with Tridacna gigas, one would expect the meat to be extracted and consumed before lugging the heavy shell to a habitation site to manufacture tools from the valves. One need only try lifting a 80 cm long individual that could weight more than 25 kg. Nearly 13 kg of shellfish were recovered from excavations at sites | and 3 and the frequency, by weight, is listed in Table 9. Overall, 56% were gastropods (mostly, Cerithidae, Turbinellidae, Turbinidae, and Neritidae) and 44% bivalves (overwhelmingly, Tridacnidae). Figure 18 compares the counts (number of identified specimens or NISP) and weights for up to 32 families from each site. Although the sample is greater at site 1 with 82.5% of total shellfish weight—and this may affect diversity measures between the sites—the relation- ship of Tridacnidae to all other taxa merits discussion. The evenness, or class (family) relative abundance, is quite different between sites. Whereas Tridacnidae makes up nearly 85% of all shell weight at site 3, itis only 33% at site 1. If the offshore islets, such as Bock, were used by small groups for short stays while capturing sea birds, reef fish, coconut crabs, and turtles, it appears that only the largest shellfish were targeted at this time. In this way, maximum yields could be acquired by short visits. Conversely, the relative abundance of shellfish families is much more even at site 1 where a more generalist collection strategy was employed. The distribution may also reflect the depletion of large bivalves during the first few centuries after atoll colonization. Hence, the greater relative emphasis on Cerithidae and a few other gastropods and bivalves. 40 Table 9. Shellfish weight for sites 1 and 3. Taxon Site 1 Site 3 Taxon Site 1 Site 3 Gastropods Trochidae 56.2 2.4 Bullidae 5.8 Trochus niloticus 27s Cardiidae 3.6 Turbo sp. 434.7 186.6 Cassis sp. 107.9 Vasum sp. 506.4 19.8 Centhium sp. 2849.9 230/) Vasum tubiferum 205.4 Centhium nodulosum 30.5 Vermetidae 7.4 Cheila equestns 1.9 Unidentified gastropod 139.4 8.7 Conus sp. (large) 31 28.8 Conus sp. (small) 229.7 20.8 Sune esse aN Cymatiidae 33.4 Bivalves Cymatium sp. 6 Arca sp. 21.2 0.6 Cypraea sp. (large) 23.4 Asaphis sp. 78.5 7.9 Cypraea sp. (small) 191.2 1.4 Barnacle 3.6 0.1 Cypraea annulus 1.4 Chama sp. 27.7 Cypraea caputserpentis 1.9 Codakia sp. Spi Cypraea moneta 5.6 Codakia punctata 14.4 Cypraeassis rufa 10.5 Codakia tigerina 727) Drupa sp. 80.5 1.9 Fragum sp. 11 0.2 Drupa grossulana alee Hippopus hippopus 65.4 605.6 Drupa morum 37.4 Isognomon sp. 0.4 Drupa ncinus 17.8 2.4 Lucinidae 5.1 Drupella sp. 2.2 Mytilidae 0.3 Lambis sp. 970.3 Penglypta sp. 2.9 Municidae 7/SU) Pinctada sp. 4.8 Nenta sp. 3:2 Pinnidae 4.5 Nenta picea 1.5 Spondylus sp. Ciel Nenta plicata 166.2 DS Tellina sp. 30.4 Nena polita 380.7 58.1 Tellina palatum 45.1 (dee Nenitidae 4.3 1.1 Tellinidae 2.3 0.4 Patellidae 5.9 Tonnidae 0.8 Sabia sp. 10.4 0.1 Tndacna sp. 1639 1290.5 Strombus sp. 33.9 Tndacna gigas 507.8 Strombus mutabilis 7.9 Tnidacna maxima 1238.1 Tectus pyramus 157.2 0.1 Tndacna squamosa 47.9 Terbridae 2.3 Veneridae 0.6 Terebra crenulata 29.1 Unidentified bivalve 72.8 1.9 to as pes Subtotal 3857.2 1908.7 t Total 10723.7 2274.9 UJAE ATOLL AS A SETTLEMENT LANDSCAPE With more than a dozen islets (total land area, 1.8 km?) surrounding a 180 km? lagoon, Ujae Atoll is one of the smaller atolls of the Marshall Islands. The majority of islets are situated along the windward coast, while the main entrance to the lagoon is through Bock Channel in the lee side. Atoll settlement patterns are predictable in a general sense, where the largest and oldest village—occupied as early as the third century—was located on the 4] eepluuo) eeplinpidaso eepuaue, gjoeweg BepIUUld BepUald aepipseg Sepiling SEpllaned aepyawie/ eepioiouoddiy SepinAW seplny aepi|Apuods eepiweyo eepuqel BepiIeWwAD @ePiqwos|s eaepiuowouBbos} SeplulleL eepiqowwesd eepisseg eepieeidAg eepiysolt eepiuoD eeploOUNW eeplUIqin, SeP|UEN eepi|jeulquny eeplujon7 SepluuED BeplusepL FIGURE 18. Counts (line) and weight (shaded) of shellfish from Ujae sites 1 and 3. 42 most substantial islet. Although a more detailed survey and subsurface testing is required for Ebbetyu and Enylamieg islets (and excavation is needed to define the precise limits of the site on Wotya), Ujae islet was, indeed, the center of population for the atoll. This is suggested further by the presence of fish traps which were used with the changing tides. The ancient Ujae village is estimated at about 25 hectares, a larger area than all the sites on the remaining islets combined. Ujae was the focus of major terrestrial production with the largest area of aroid (presumably Cyrtosperma chamissonis) pit cultivation and thickest Ghyben-Herzberg fresh water lens. Prehistoric settlement may, in fact, mirror the extent of the modern village with its nearly 500 inhabitants. At least 200-300 residents seems likely, although admittedly, this is a best guess. While Enylamieg probably supported the next largest village (and probably had aroid pit cultivations), Wotya and Ebbetyu may have sus- tained on the order of 10-25 full-time residents, yet no evidence of pit agriculture was seen there. The smallest islets were not large enough to retain groundwater and, consequently, could not have sustained permanent habitation. Before human colonization and settlement of Ujae Atoll, all of the forested land prob- ably supported dense colonies of sea birds and ocean side beaches provided nesting sites for the green sea turtle. With the build-up of human populations on the largest islets of Ujae and Enylamieg, prize resources like coconut crab and sea turtles were depleted and sea bird colonies may have shifted to the smaller offshore islets (such as Bock, Bokerok, and Wotya) or, perhaps, birds on these smaller islets represent remnant populations of the atoll. Indeed, today, sea bird colonies, turtles, and coconut crabs are only found on the smaller islets. From the main islet of Ujae, it is about 12 nautical miles to the channel and the ocean side reef, making trolling sorties for pelagic species a time-consuming and labor-costly prac- tice. During our stay, traditional-style sailing canoes would travel northwest inside the la- goon for bottom fishing, staying within about 5 km of the main village. If the channel was used to gain access to the pelagic fisheries during prehistoric times, it is likely that Bock islet would have supported temporary encampments of small groups before and after fish- ing in the more protected lee of the atoll. Much of the broad reef flats north of Ujae (see Fig. 2) are exposed at low tide when octopus and shellfish (especially Cerithium nodulosum) are collected. Stone fish traps and weirs were built on the reef flats for channelling and captur- ing fish and were situated to take advantage of local topographic conditions such as natural alignments of raised reef and, on the west coast of Ujae, they were constructed with their long axis parallel to the prevailing out-going current which trapped fish within the con- stricted end. The marine habitats remained relatively stable over the two millennia of human occupa- tion, yet, during this time, the effects of human predation left its signature. Bones of turtle were quite rare, suggesting depletion of previously more abundant stocks. Because the earliest few centuries of human colonization of Ujae Atoll have not been documented, it is likely that turtle bones were more common in the earliest deposits, a situation noted for other islands (Weisler 1995). The same can be said in reference to the lack of bones of extinct birds—usually the temporal domain of the first few hundred years of human occu- pation, especially on small atolls. Considering the marine shellfish, the earliest cultural deposits should contain a relatively high abundance of large individuals, especially of taxa such as Tridacnidae and Strombidae (Lambis spp.). Rather, the main ancient village at Ujae 43 islet shares quantities of Tridacna similar to those of Cerithium—a species with a much lower meat weight per specimen and a relatively light shell. This suggests that Cerithium was an important food shellfish, perhaps after Tridacna stocks were reduced. Lambis spp. are quite rare in the Ujae assemblage suggesting this high meat shellfish was depleted early on. The artifact assemblage is typical for an atoll setting with great use made of large shell- fish for fashioning a range of tools and ornaments. Although it is tempting to argue for some manner of contact between the Ujae population and the volcanic island of Kosrae, or even more distant islands on the continental side of the Andesite Line, the geological lo- cales of exotic rocks or manuports may simply signal the direction of the dominant ocean swells that cast drift trees ashore with foreign rocks entangled in their roots. One cannot, however, rule out the possibility of long distance voyaging, so well documented for other parts of the Pacific (e.g., Weisler 1997). The Ujae archaeological study has been the first phase of a longer-term investigation of atolls situated along the rainfall gradient of the Marshall Islands. The differing rainfall regimes and the variability of atoll size, shape, and number of islets should be expressed in prehistoric settlement patterns whose diachronic study will illuminate how humans colo- nized, adapted, and transformed the most precarious of Pacific landscapes. ACKNOWLEDGMENTS Permission to work on Ujae Atoll was granted by Irooj Michael Kabua who also pro- vided wonderful accommodations. Dennis Alessio and family were instrumental for logis- tics on Ujae and for facilitating on-island amenities. The field work, supported in part by the U. S. National Park Service through its grants-in-aid program, was completed while I was Chief Archaeologist for the Republic of the Marshall Islands and I thank Carmen Bigler for additional aid from the Historic Preservation Office. The research committee of the University of Otago is thanked for funding additional lab work. The efforts of Dave Steadman and Mike Green are appreciated for identifying the bird bones and human remains, respec- tively. Artifact illustrations are by Wren and Les O’Neill (who also prepared the other fig- ures). And at the Bernice P. Bishop Museum Archives, Betty Lou Kam, as always, provided much appreciated help. 44 REFERENCES Abo, T., Bender, B. W., Capelle, A. and DeBrum, T. 1976. Marshallese-English Dictionary. Honolulu: University of Hawaii Press. Allen, M. S. 1992. Temporal variation in Polynesian fishing strategies: the Southern Cook Islands in regional perspective. Asian Perspectives 31:183-204. Anderson, A. 1997. Uniformity and regional variation in marine fish catches from prehis- toric New Zealand. Asian Perspectives 36:1-26. Beardsley, F. R. 1994. Archaeological Investigations on Kwajalein Atoll, Marshall Islands. International Archaeological Research Institute, Inc., Honolulu. Bellwood, D. R. 1994. A Phylogenetic Study of the Parrotfishes Family Scaridae (Pisces: Labroidei), with a Revision of Genera. Records of the Australian Museum, Supplement 20. Bellwood, D. R. and Choat, J. H. 1990. A functional analysis of grazing in parrotfishes (family Scaridae): the ecological implications. Environmental Biology of Fishes 28:189- 214. Brooke, M. de L. 1995. Seasonality and numbers of green turtles Chelonia mydas nesting on the Pitcairn Islands. Biological Journal of the Linnean Society 56:325-327. Bryan, Jr., E. H. 1971. Guide to Place Names in the Trust Territory of the Pacific Islands (the Marshall, Caroline and Mariana Is.). Honolulu: Pacific Science Information Center, Bernice P. Bishop Museum. Butler, V. L. 1994. Fish feeding behaviour and fish capture: the case for variation in Lapita fishing strategies. Archaeology in Oceania 29:81-90. Carucci, J. 1992. Cultural and natural patterning in prehistoric marine foodshell from Palau, Micronesia. PhD thesis, Southern Illinois University at Carbondale. Dean, J. S. 1978. Independent dating in archaeological analysis. Advances in Archaeologi- cal Method and Theory 1:223-255. Dye, T. 1983. Fish and fishing on Niuatoputapu. Oceans 53:247-271. Dye, T. 1987. Archaeological survey and test excavations on Arno Atoll, Marshall Islands. In T. Dye (ed.), Marshall Islands Archaeology, pp. 271-399. Pacific Anthropological Records 38. Honolulu: Bernice P. Bishop Museum. Emory, K. P. 1939. Archaeology of Mangareva and Neighboring Atolls. Bernice P. Bishop Museum Bulletin 163. 45 Emory, K. P., Bonk, W. J., and Sinoto, Y. H. 1959. Hawaiian Archaeology: Fishhooks. Bernice P. Bishop Museum Special Publication 47. Finsch, O. 1893. Ethnologische Erfahrungen und Belegstiicke aus des Sitidsee. Dritte Abtheilung: Mikronesien (II. Marshall-Archipel; II. Carolinen [1. Kuschai, 2. Ponapé]). Annalen des K. K. Naturhistorischen Hofmuseums pp. 119-182. Fosberg, F. R. 1955. Northern Marshall Islands Expedition, 1951-1952. Narrative. Atoll Research Bulletin 38. Washington, D. C.: National Academy of Sciences. Fosberg, F. R. 1984. Phytogeographic comparison of Polynesia and Micronesia. Bernice P. Bishop Museum Special Publication 72:33-44. Hatheway, W. H. 1953. The Land Vegetation of Arno Atoll, Marshall Islands. Atoll Re- search Bulletin 16. Washington, D. C.: Pacific Science Board, National Academy of Sci- ences. Hastorf, C. A. and Popper, V. S. eds. 1988. Current Paleoethnobotany: Analytical Methods and Cultural Interpretations of Archaeological Plant Remains. Chicago: University of Chi- cago Press. Hooper, P. R., Johnson, D. M. and Conrey, R. M. 1993. Major and trace element analyses of rocks and minerals by automated x-ray spectrometry. Manuscript on file, Department of Geology, Washington State University, Pullman. James, H. F. and Olson, S. L. 1991. Descriptions of Thirty-two New Species of Birds from the Hawaiian Islands: Part II. Passeriformes. Ornithological Monographs No. 46. Wash- ington, D. C.: American Ornithologicalists’ Union. Johannes, R. E. 1981. Words of the Lagoon: Fishing and Marine Lore in the Palau District of Micronesia. Berkeley: University of California Press. Kikuchi, W. K. 1976. Prehistoric Hawaiian fishponds. Science 193:295-299. Kirch, P. V. 1988. Long-distance exchange and island colonization: the Lapita case. Nor- wegian Archaeological Review 21:103-117. Kirch, P. V. 1979. Marine Exploitation in Prehistoric Hawai‘i: Archaeological Investiga- tions at Kalahuipua’a, Hawai‘i Island. Pacific Anthropological Records 29. Honolulu: Bernice P. Bishop Museum. Kirch, P. V. and Dye, T. 1979. Ethnoarchaeology and the development of Polynesian fishing strategies. Journal of the Polynesian Society 88:53-76. Kirch, P. V. and Weisler, M. I. 1994. Archaeology in the Pacific islands: an appraisal of recent research. Journal of Archaeological Research 2:285-328. 46 Kirch, P. V. and Yen, D. E. 1982. Tikopia, The Prehistory and Ecology of a Polynesian Outlier. Bernice P. Bishop Museum Bulletin 238. Leach, B. F. and Anderson, A. 1979. The role of labrid fish in prehistoric economies in New Zealand. Journal of Archaeological Science 6:1-15. Leach, B. F. and Boocock, A. S. 1993. Prehistoric Fish Catches in New Zealand. Oxford: BAR International Series 584. Leach, B. F., Davidson, J. and Athens, J. S. 1996. Mass harvesting of fish in the waterways of Nan Nadol [sic], Pohnpei, Micronesia. In Davidson, J., G. Irwin, F. Leach, A. Pawley, and D. Brown (eds.), Oceanic Culture History: Essays in Honour of Roger Green, pp. 319- 341. Dunedin: New Zealand Journal of Archaeology Special Publication. Leach, B. F., Davidson, J., Horwood, M. and Ottino, P. 1997. The fisherman of Anapua rock shelter, Ua Pou, Marquesas. Asian Perspectives 36:51-66. Leach, B. F., Intoh, M. and Smith, I. W. G. 1984. Fishing, turtle hunting, and mammal exploitation at Fa‘ahia, Huahine, French Polynesia. Journal de la Société des Océanistes 40:183-197. Mason, L. 1947. The Economic Organization of the Marshall Islands. Honolulu, U. S. Commercial Co. Mason, R. D., Peterson, M. L. and Tiffany, J. A. 1998. Weighing vs. counting: measurement reliability and the California school of midden analysis. American Antiquity 63:303-324. Myers, R. F. 1991. Micronesian Reef Fishes. (2nd ed.). Barrigada, Guam: Coral Graphics. Neal, M. C. 1965. In Gardens of Hawaii. Bernice P. Bishop Museum Special Publication 50. Office of Planning and Statistics, 1988. Census of Population and Housing 1988, Final Report. Majuro, Marshall Islands: Office of Planning and Statistics. Randall, J. E. 1996. Shore Fishes of Hawai‘i. Vida, Oregon: Natural World Press. Riley, T. J. 1987. Archaeological survey and testing, Majuro Atoll, Marshall Islands. In T. Dye (ed.), Marshall Islands Archaeology, pp. 169-270. Pacific Anthropological Records 38. Honolulu: Bernice P. Bishop Museum. Rolett, B. 1989. Hanamiai: changing subsistence and ecology in the prehistory of Tahuata (Marquesas Islands, French Polynesia). PhD thesis. New Haven: Yale University. Rosendahl, P. H. 1987. Report 1: Archaeology in eastern Micronesia: reconnaissance sur- vey in the Marshall Islands. In T. Dye (ed.) Marshall Islands Archaeology, pp. 17-168. Pacific Anthropological Records 38. Honolulu: Bernice P. Bishop Museum. 47 Sachet, M.-H. 1955. Pumice and Other Extraneous Volcanic Materials on Coral Atolls. Atoll Research Bulletin 37, Washington, D. C. Shun, K. and Athens, J. S. 1990. Archaeological investigations on Kwajalein Atoll, Marshall Islands, Micronesia. Micronesica Supplement 2:231-240. Sinoto, Y. H. 1959. Solving a Polynesian fishhook riddle. Journal of the Polynesian Society 68:23-28. Sinoto, Y. H. 1962. Chronology of Hawaiian fishhooks. Journal of the Polynesian Society 71:162-166. Somerville-Ryan, G. 1998. The taphonomy of a Marshall Islands’ shell midden. M. A. thesis, University of Otago. Spennemann, D. H.R. and Ambrose, W. R. 1997. Floating obsidian and its implications for the interpretation of Pacific prehistory. Antiquity 71:188-193. Steadman, D. W. 1989. Extinction of birds in Eastern Polynesia: a review of the record, and comparisons with other Pacific island groups. Journal of Archaeological Science 16:177- 205. Steadman, D. W. 1995. Prehistoric extinctions of Pacific island birds: biodiversity meets zooarchaeology. Science 267:1123-1131. Stoddart, D. R. 1992. Biogeography of the tropical Pacific. Pacific Science 46:276-293. Stone, B. C. 1960. The wild and cultivated Pandanus of the Marshall Islands. M. A. thesis. Honolulu: University of Hawai‘i. Streck, C. S. 1990. Prehistoric settlement in eastern Micronesia: archaeology of Bikini Atoll, Republic of the Marshall Islands. Micronesica Supplement 2:247-260. Stuiver, M. and Reimer, P. J. 1993. Extended !4C data base and revised 3.0 '4C age calibra- tion program. Radiocarbon 35:215-230. Summers, C. C. 1964. Hawaiian Fishponds. Bernice P. Bishop Museum Special Publica- tion 52. Thompson, S. 1982. Cyrtosperma chamissonis (Araceae): ecology, distribution, and eco- nomic importance in the South Pacific. Journal d’Agriculture Tropicale et de Botanique Applique 29:185-203. Walter, R. 1998. Anai’o: The Archaeology of a Fourteenth Century Polynesian Community in the Cook Islands. New Zealand Archaeological Association Monograph 22. 48 Weisler, M. I. 1995. Henderson Island prehistory: colonization and extinction on a remote Polynesian island. Biological Journal of the Linnean Society 56:377-404. Weisler, M. I. ed. 1997. Prehistoric Long-distance Interaction in Oceania: An Interdiscipli- nary Approach. New Zealand Archaeological Association Monograph 21. Weisler, M. I. 1999. Prehistoric farmers of the Marshall Islands. Discovering Archaeology 1(1):66-69. Weisler, M. I. In press. The antiquity of aroid pit agriculture and significance of buried A horizons on Pacific atolls. Geoarchaeology. Weisler, M. I. and Gargett, R. H. 1993. Pacific island avian extinctions: the taphonomy of human predation. Archaeology in Oceania 28:85-93. Weisler, M. I., Lum, J. K., Collins, S. L. and Kimoto, W. S. In press. Status, health, and ancestry of a late prehistoric burial from Kwajalein Atoll, Marshall Islands. Micronesica. Weisler, M. I. and Murakami, G. M. 1991. The use of Mountain-apple (Syzygium malaccense) in a prehistoric Hawaiian domestic structure. Economic Botany 45:282-285. Weisler, M. I. and Walter, R. 1999. Late prehistoric fishing adaptations at Kawakiu Nui, West Moloka‘i. Hawaiian Archaeology 8. Widdicombe, H. 1997. The cutting edge: a technological study of adzes from Ebon, Maloelap and Ujae atolls, Marshall Islands. M. A. thesis, University of Otago. Wiens, H. J. 1959. Atoll development and morphology. Annals of the Association of Ameri- can Geographers 49:31-54. Williamson, I. and Sabath, M. D. 1982. Island population, land area, and climate: a case study of the Marshall Islands. Human Ecology 10:71-84. Wragg, G. M. and Weisler, M. I. 1994. Extinctions and new records of birds from Henderson Island, Pitcairn Group, South Pacific Ocean. Notornis 41:61-70. 49 APPENDIX 1: STRATIGRAPHIC DESCRIPTIONS The following four profiles from site | record the stratigraphic variability along the 420 m excavation transect oriented 180° from the middle of the lagoon shore of Ujae islet. Unit 6, situated 14 m inland from the bottom of the vegetation line at the beach, begins the transect; unit 5 is 94 m inland; unit 2 is 218 m inland; and unit 4 is 420 m inland (see Fig. 7). The stratigraphy for units 3, 4, 7, and 8 are nearly identical in layer characteristics. Profile Description of Site 1, Unit 6 (Fig. 12) I Historic cultural layer and loose sand; 0-3 cm bs; average thickness is 3 cm; sharp, smooth boundary; 1OYR 6/1 (gray); medium sand texture; structureless (grade), very fine (size), single grain (form) structure; loose (dry), loose (moist), and non-sticky (wet) consistence; non-plastic; no cementation; abundant roots and pores. II Historic cultural layer and gravelly sand; 3-6 cm bs; average thickness is 3 cm; sharp, smooth boundary; 1OYR 5/1 (gray); gravelly sand texture consisting of medium sand with water-rounded coral pebbles; weak, very fine, single grain structure; loose, loose, and non-sticky consistence; non-plastic; no cementation; abundant roots and pores. III Prehistoric to protohistoric cultural layer dated to 100.6 + 0.6% modern (Beta-7958 1) with dense fish bone and few food shellfish; 6-37 cm bs; average thickness is 31 cm; gradual, wavy boundary; 1O0YR 3/1 (very dark gray); medium sand texture; structureless, very fine, single grain structure; loose, loose, and non-sticky consistence; non-plastic; no cementation; abundant roots and pores. IV Culturally sterile subsoil; 37-50 cm bs; average thickness is 13 cm, but bottom was not reached; boundary could not be determined; 1OYR 6/3 (pale brown) with 10YR6/1 (light gray) mottles; medium sand; structureless, very fine, single grain structure; loose, loose, and non-sticky consistence; non-plastic; no cementation; abundant roots and pores. Profile Description of Site 1, Unit 5 (Fig. 11) I Contemporary water-rounded coral pavement; 0-5 cm bs; average 5 cm thick; very abrupt, smooth boundary; 5Y 7/1 (light gray); gravel texture; structureless (grade), very coarse (size), single grain (form); loose (dry), loose (moist), non-sticky (wet) consistence; non- plastic; no cementation; no roots and pores. II Contemporary “floor” or base of coral pavement; 5-7 cm bs; average 2 cm thick; abrupt, smooth boundary; 2.5Y 3/3 (very dark gray); very fine to coarse sand with small gravel texture; structureless, fine, crumb; slightly hard, friable, slightly sticky; slightly plastic; no cementation; no roots or pores. III Prehistoric cultural layer; 7-37 cm bs; average 30 cm thick; clear, wavy boundary; 2.5Y 2/0 (black); coarse sand texture; weak, fine, crumb structure; weakly coherent, friable, slightly plastic; slightly plastic; no cementation; fine, very few roots and pores. IV Prehistoric cultural layer dated to 560 + 70 BP (Beta-76018); 37-84 cm bs; average 47 cm thick; clear, wavy boundary; 10YR 4/1 (dark gray); coarse sand texture; structureless, 50 fine, single grain; loose, loose, non-sticky; non-plastic; no cementation; no roots or pores. : Sterile subsoil; 84-90+ cm bs; greater than 6 thick; boundary not visible; 5Y 7/3 (pale yellow); coarse sand texture; structureless, fine, single grain; loose, loose, non-sticky; non-plastic; no cementation; no roots or pores. Profile Description of Site 1, Unit 2 (Fig. 11) I II Prehistoric cultural layer; 0-103 cm bs; average thickness is 103 cm; gradual, smooth boundary; 7.5YR 2/0 (black); gravelly sand texture; moderate (grade), medium (size), crumb (form) structure; loose (dry), very friable (moist), slightly sticky (wet) consist- ence; slightly plastic; no cementation, abundant, many roots and pores. Prehistoric cultural layer dating to 1660 + 60 BP (Beta-74845) grading to sterile, moist, compact gravelly sand; 103-136 cm bs; average thickness is 33 cm; boundary not vis- ible; 7.5YR 3/0 (very dark gray); gravelly sand; moderate, medium, crumb; weakly coherent, very friable, slightly sticky; slightly plastic; no cementation, abundant, many roots and pores. Profile Description of Site 1, Unit 4 (Fig. 11) I I Sparse cultural layer with dense gravel; 0-40 cm bs; average thickness is 40 cm; smooth, clear boundary; 2.5YR 2.5/0 (black); gravelly sand texture; weak (grade), medium (size), crumb (form) structure; weakly coherent (dry), very friable (moist), slightly sticky (wet) consistence; slightly plastic; no cementation, abundant, many coconut roots and pores. Excavated with rock hammer. Very sparse cultural; 40-60 cm bs; average thickness is 17 cm; smooth, clear boundary; 1O0YR 4/1 (dark gray); gravelly sand texture; weak, medium, crumb structure; weakly coherent, very friable, slightly sticky consistence; slightly plastic; no cementation, abun- dant, many coconut roots and pores. Less compact than layer I, but excavated with rock hammer. If] Culturally sterile; 50-60 cm bs; average thickness is 10 cm; boundary could not be determined; 5YR 7/3 (pink); medium to coarse sand texture; structureless, fine, single grain; loose, loose, non-sticky; non-plastic; no cementation, abundant, many coconut roots and pores. Profile Description of Site 3, Unit 1 (Fig. 13) I Historic coral pavement, glass, and metal fragments; 0-7 cm bs; average thickness is 7 cm; abrupt, smooth boundary; 7.5YR 2/0 (black); medium to coarse sand texture with dense, water-rounded coral gravel; weak (grade), fine (size), and single grain (form) structure; loose (dry), loose (moist), and non-sticky (wet) consistence; non-plastic; no cementation; abundant roots and pores. Mixed historic (7-20 cm bs) and prehistoric (20-30 cm bs) cultural deposits (separated approximately by the dashed line in Fig. 13); total average thickness of 28 cm; abrupt, Sil wavy boundary; 7.5YR 2/0 (black); medium to coarse sand with dense, water-rounded coral gravel; weak, fine, and single grain structure; loose, loose, and non-sticky consist- ence; slightly plastic; no cementation; abundant roots and pores. III Main prehistoric cultural layer (35-96 cm bs) with dense ovens; radiocarbon dated to 30 + 50 BP (Beta-79583) to 200 + 60 BP (Beta-76019); average thickness is 61 cm; clear, wavy boundary; 2.5Y 2/0 (black) with a sterile, coarse sand pocket of 1OYR 7/2 (light gray); medium to coarse sand with dense cobbles (oven stones); structureless, fine, sin- gle grain; loose, loose, non-sticky consistence; non-plastic; no cementation; abundant, many roots. IV Sparse prehistoric cultural layer (96-118 cm bs); average thickness is 22 cm; clear, wavy boundary; 10YR 3/1 (very dark gray); medium to coarse sand with water-rounded coral gravel; structureless, fine, single grain; loose, loose, non-sticky consistence; non- plastic; no cementation; abundant, many roots. V_ Culturally sterile subsoil (118-140 cm bs); average thickness is 52+ cm as bottom was not reached; boundary could not be determined; 10YR 7/2 (light gray); coarse sand and subrounded coral cobbles, 12-20 cm in maximum length; structureless, fine, single grain; loose, loose, non-sticky consistence; non-plastic; no cementation; abundant, many roots. a re antigh cit” but, qyeon.ct paaieane cml OS £ oa dnt a" Wir (ott bith 3800! daK jiu? fey he nie ar MN HN r Lreverservaey cote citiamleg | qi a of ha) oO . nat 4 is iis wD )ii 7 q | nol ee Patel) Ga0¢ ; wubnuii yy | it , . “Tach si 1) Thy Mined Histmie (7-20 or) ar preinete CRD Aa ey @ ia 0 > deprosivasmiy y the dubes lag oe ems 2 ATOLL RESEARCH BULLETIN NO. 461 REPORT ON FISH COLLECTIONS FROM THE PITCAIRN ISLANDS BY JOHN E. RANDALL ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. AUGUST 1999 Figure 1. Oeno Atoll (aerial photograph by Gerald R. Allen, 1969). Young's Rocks ADAMSTOWN PITCAIRN ISLAND Tekewtema ial Figure 2. Map of Pitcairn Island (modified from A Guide to Pitcairn, British South Pacific Office, Suva, 1970). REPORT ON FISH COLLECTIONS FROM THE PITCAIRN ISLANDS BY JOHN E. RANDALL! ABSTRACT A total of 348 species of marine fishes are recorded from the four Pitcairn Islands in southeastern Oceania: Pitcairn, Henderson, and the atolls Oeno and Ducie. Nearly all of the species listed are from collections made by the author and associates in 1970-71 and deposited in the Bernice P. Bishop Museum, Honolulu. Thirty-three of these were new species when they were collected but have since been described. Twenty-six species are listed only by genus, 15 of which appear to be undescribed and are under study; the remaining 11 are unidentified. Five species of fishes are presently known only from the Pitcairn Islands: Sargocentron megalops, Hemitaurichthys multispinosus, Ammodytes sp., Enneapterygius ornatus, and Alticus sp. Of the 335 species from the Pitcairn Islands that are shore fishes (the other 13 being regarded as pelagic), 284 are tropical species that are wide-ranging in the central and western Pacific, many of which extend their distribution into the Indian Ocean. Thirty-six of the Pitcairn fishes occur only in the Southern Hemisphere south of latitude 14° S; 21 of these are found only south of 20°S. Some of the latter are broadly distributed at islands across the southern subtropical zone. Twelve species of Pitcairn fishes are antitropical or antiequatorial in distribution based on present knowledge. INTRODUCTION The islands of the Pitcairn Group in southeastern Oceania consist of Pitcairn (25°4’S, 130°6’W), Henderson, and the atolls of Oeno and Ducie. Only Pitcairn is inhabited. Lying just south of the Tropic of Capricorn, these four islands are near the southern limit of distribution for tropical Indo-Pacific species of fishes. The nearest island is Temoe Atoll of the Tuamotu Archipelago, about 400 km west of Oeno. To the east there is only Easter Island, more than 1700 km away. 1Bishop Museum, 1525 Bernice St., Honolulu, Hawaii 96817-2704 Manuscript received 23 February 1999; revised 13 July 1999 For a month beginning 18 December 1970, the author visited the four Pitcairn Islands in the 30-m schooner “Westward” with support of a grant from the National Geographic Society for fish collecting (Randall, 1973, 1978). Also aboard were the late Dr. Harald A. Rehder, malacologist of the Smithsonian Institution, Dr. Yoshihiko Sinoto, then Chairman of Anthropology of Bishop Museum, the late Dennis M. Devaney, then Chairman of Marine Invertebrate Zoology of Bishop Museum, Guy S. Haywood, M.D., the late Captain W. Roger Gray, and a crew of nine, most of whom were also divers. In addition to the documentation of the fish fauna of these remote islands, there was interest to know the composition of the fish fauna, i.e. how many are wide-ranging tropical Indo-Pacific species; how many are shared with Easter Island to the east (fishes reported by Randall and Cea Egafia, 1984; DiSalvo et al., 1988); how many with Rapa (Randall et al. , 1990) and other southern subtropical islands to the west; and whether any are endemic to the Pitcairn Islands. After spending a week at Mangareva and a day at Temoe Atoll in the southeastern Tuamotus, the “Westward” made its first landfall in the Pitcairn Islands at Oeno (23°55’S, 130°45’W) (Fig. 1), about 4 km in diameter, with a shallow lagoon, no pass, and no safe anchorage. Two days were spent collecting fishes with the ichthyocide rotenone (two stations), hook and line, spearing, and powerhead (explosive-tipped spear) The vessel arrived in Pitcairn on 20 December and anchored off Bounty Bay. The island (Fig. 2) is small, about 3 km long, of rugged terrain, with no permanent streams and no sandy beaches. A small photographic laboratory was established at “The Edge” above Bounty Bay (Fig. 3), and some fishes that were brought frozen from Temoe and Oeno were photographed. From 21 December to 10 January, and again from 18 January to 21 January, a total of 31 fish-collecting stations were carried out at Pitcairn; ten of these stations were with rotenone and the rest with spears and quinaldine. In addition, some fishes were taken by hook and line, and a few specimens were provided by Pitcairn fishermen. After the first two days, high winds and heavy seas often made collecting difficult. The rough seas prevented the Pitcairn islanders from deep hand-line fishing, as planned, so all our fish specimens from Pitcairn came from our maximum diving depth of 55 m or less. The surf along the shore also limited inshore collecting. However, several large tidepools, such as the nearly landlocked one at Down St. Paul (Fig. 4) were very productive. The shallower diving revealed a bottom usually dominated by large boulders (Fig. 5), heavily covered with brown algae (mainly Sargassum coriifolium and the branching form of Lobophora variegata; in the surge zone, primarily Turbinaria). In some areas, such as off Gudgeon Harbor on the south side, there 3 are broad patches of soft sand deeply etched with ripple marks. At depths greater than about 15 m, the bottom is mainly sand with small rocky areas of low profile. With information from the islanders and by towing divers, we found some well-developed coral reefs. One off Adamstown at a depth of 22-30 m covers an area of about 2 square kilometers, about half consisting of live coral (the colonies less than about 70 cm in diameter). Another off the north shore known by the fishermen as “The Bear”, rises about 9 m above the substratum at 44.5 m; it is unusual in being almost entirely covered by live coral. The collecting there provided some very valuable fishes such as Sargocentron ensifer, Genicanthus spinus, and Hemitaurichthys multispinosus. Despite much searching, no steep dropoffs were found off Pitcairn. The sea temperature averaged about RAG On 11 January we sailed to Henderson Island (24°21’S, 128°19’W) to take 14 islanders and their 28-foot boat to collect miro wood for their carving and to leave Dr. Sinoto and his son Aki for archaeological field work while we went on to Ducie. Henderson is a limestone island, the flattened top with a maximum height of about 30m. Cliffs of about 15 m in height drop abruptly into the sea along the western and southern shores. There is a long sandy beach on the north side, and most of our fish collecting activity was off this beach. Spearing and quinaldine collections were made for a half day on 11 January. Two rotenone stations were carried out during a day and a half after return from the visit to Ducie, one in 49 m resulting in the only specimen of a new species of Sargocentron. Sharks were more troublesome at Henderson than the other islands, and two gray reef sharks were powerheaded by divers during collecting activities; four others to a length of 1.7 m were caught by hook and line. R.A. Irving and J. Jameson were on Henderson Island from October to December, 1991 as part of the Sir Peter Scott Commemorative Expedition to the Pitcairn Islands. Their observations and photographic records of fishes at the island were added to the author's collections and reported as a checklist of 183 species (Irving et al., 1995). Two and one-half days were spent at Ducie (24°40’S, 124°47’W), the easternmost atoll in Oceania, 470 km east of Pitcairn. An account of the history, physiography, and biota of the atoll was published by Rehder and Randall (1975); it includes a preliminary checklist of 110 species of fishes that were collected and another 27 species as sight records. Among the fishes, the southern subtropical species of butterflyfish, Chaetodon flavirostris (Fig. 6). The following is a list of the species of fishes collected and observed at the Pitcairn Islands. SL stands for standard length, FL for fork length, TL for total length, and PCL for precaudal length (of sharks). The number of specimens reported for the species does not necessarily reflect the relative abundance of the 4 species. In fact, in some cases, just the reverse, as special effort was often made to collect more specimens of rare or unidentified species. Also, fewer specimens were kept of the larger species because of limitation of the supply of formalin, containers, and space on the vessel. During the 7-months cruise on “Westward”, fishes and other marine organisms were also collected from the Society Islands, Tuamotu Archipelago, Rapa, Austral Islands, Cook Islands, and the Marquesas. Illustrations are given for 32 species of fishes from photographs of specimens collected in the Pitcairn Islands. Species chosen for the illustrations are primarily ones that have not been well illustrated previously or ones for which the photographs help in confirming the identification. Thirty-three species of fishes from the Pitcairn Islands, undescribed at the time of collection, have been named. All but three of these were represented by specimens from other localities when they were described. The Bishop Museum type specimens of these new species are indicated by parenthetical remarks in the listing of specimens below. Twenty-six species of Pitcairn fishes are listed by generic name only. Fifteen of these appear to be undescribed and are under study by ichthyologists. The remaining 11 are unidentified due to need for more systematic study of the genera in question, poor condition or juvenile status of specimens, two records from sightings, and in one case to the loss of the only specimen while on loan. The following publications included descriptions of new species of Pitcairn fishes based at least in part on specimens from the Bishop Museum’s collections: Allen and Randall (1972), Cressey and Randall (1978), Fricke (1997), Greenfield (1974), Herald and Randall (1974), Lavenberg (1992), McCosker and Randall (1977), H. Randall and Allen (1977), Randall (1972), Randall (1975a), Randall (1975b), Randall (1979), Randall (1980), Randall (1998), Randall and Baldwin (1997), Randall and Caldwell (1973), Randall and Choat (1980), Randall and Edwards (1984), Randall and McCosker (1992), Randall and Pyle (1989), Randall and Randall (1981), Randall and Smith (1988), and Randall (1999). Paulay (1989) published a report on the corals, mollusks, and echinoderms of the Pitcairn Islands. LIST OF SPECIES AND SPECIMENS CARCHARHINIDAE (REQUIEM SHARKS) Carcharhinus amblyrhynchos (Bleeker, 1856). BPBM 17073, 680 mm PCL, BPBM 17108, 840 mm PCL, Henderson. Carcharhinus galapagensis (Snodgrass and Heller, 1905). BPBM 12573, 1460 5 mm PCL, Pitcairn; also reported from Ducie (Rehder and Randall, 1975). Carcharhinus melanopterus (Quoy and Gaimard, 1824). Henderson (Irving et al., 1995). Galeocerdo cuvier (Peron and Lesueur, 1822). Two tiger sharks have been reported from Pitcairn, one of 4 m (Pitcairn Miscellany, 1980, vol. 22, no. 1), and one of 2.5 m (Pitcairn Miscellany, 1985, vol. 27, no. 9). Triaenodon obesus (Riippell, 1835). Five collected at Oeno, 1105-1238 mm PCL (two shown in Fig. 7), but none was retained; observed at the other three islands. MYLIOBATIDAE (EAGLE RAYS) Aetobatis narinari (Euphrasen, 1790). Not observed during the expedition on “Westward”, but Stephen Christian sent a photograph of a specimen caught off the jetty at Bounty Bay, Pitcairn in May, 1974 that measured 5 feet in disc width and was estimated to weigh more than 100 pounds. Another ray of this species was harpooned by Christian at Pitcairn in 1983 and reported in Pitcairn Miscellany, 1983, vol. 25, no. 3: MORINGUIDAE (WORM EELS) Moringua sp. Six lots of eels of this genus were collected at the Pitcairn Islands. No attempt at this time has been made to identify the specimens to species. Difference in morphology of males and females, including more vertebrae in males, for the Hawaiian species of Moringua (Gosline and Strasburg, 1956), and the conclusions of Castle (1968) in his valuable preliminary revision of the genus indicate the need for a complete revision of this difficult genus. CHLOPSIDAE (FALSE MORAYS) Kaupichthys diodontus Schultz, 1943. BPBM 16615, 99 mm TL, Oeno; BPBM 16890, 76 mm TL, Pitcairn; BPBM 17006, 124 mm TL, Pitcairn; BPBM 17031, 106 mm TL, Pitcairn. Identifications by Kenneth A. Tighe. CONGRIDAE (CONGER EELS) Ariosoma sp. BPBM 16621, 181 mm TL, Oeno. Under study by David G. Smith who reports the count of predorsal/preanal/total vertebrae as 7/56/151. This may eventually be shown to be within the range of A. marginatum (Vaillant and Sauvage). Conger cinereus cinereus Rippell, 1828. BPBM 16721, 252 mm TL, Pitcairn; BPBM 17007, 135 mm TL, Pitcairn. OPHICHTHIDAE (SNAKE EELS) Apterichtus sp. BPBM 11842, 347 mm TL, Pitcairn; BPBM 16506, 324 mm TL, Oeno; BPBM 16896, 2: 293-385 mm TL, Pitcairn. Initially identified as A. flavicaudus (Snyder), but now believed to be undescribed (Fig. 15). Ichthyapus vulturis (Weber and de Beaufort, 1916). BPBM 16467, 217 mm TL, Oeno; BPBM 16527, 112 mm TL, Oeno. Leiuranus semicinctus (Lay and Bennett, 1839). BPBM 16725, 215 mm TL, Pitcairn; BPBM 16988, 165 mm TL, Pitcairn. Muraenichthys gymnotus Bleeker, 1857. BPBM 17023, 4: 110-165 mm TL, Pitcairn. Muraenichthys laticaudata (Ogilby, 1897). BPBM 12263, 2: 103-106 mm TL, Ducie; BPBM 16487, 170 mm TL, Oeno; BPBM 16507, 236 mm TL, Oeno; BPBM 16625, 215 mm TL, Oeno; BPBM 16732, 3: 107-135 mm TL, Pitcairn; BPBM 16971, 2: 112-133 mm TL, Pitcairn; BPBM 16980, 147 mm TL, Pitcairn; BPBM 17093, 11: 91-223 mm TL, Henderson. (Fig. 16). Myrichthys maculosus (Cuvier, 1816). BPBM 16456, 500 mm TL, Pitcairn. MORAY EELS (MURAENIDAE) Anarchias spp. Seven lots of small morays of this genus from the Pitcairn Islands are under study by David G. Smith and Erling Holm. Six are provisionally identified as A. seychellensis Smith; the seventh is an undescribed species similar to A. cantonensis (Schultz) but with a low count of about 100 vertebrae. Enchelynassa canina (Quoy and Gaimard, 1824). Henderson (Irving et al., 1995). Gymnothorax australicola Lavenberg, 1992. BPBM 12262, 257 mm TL, Ducie (misidentified as G. panamensis by Rehder and Randall, 1975); BPBM 16844, 230 mm TL (paratype), Pitcairn; BPBM 16877, 5: 180-305 mm TL (paratypes), Pitcairn; BPBM 16975, 3: 115-190 mm TL, Pitcairn; BPBM 16991, 2: 145-155 mm TL, BPBM 17014, 290 mm TL (paratype), Pitcairn; BPBM 17036, 3: 130-258 mm TL, Pitcairn. Gymnothorax buroensis (Bleeker, 1857). BPBM 16690, 196 mm TL, Pitcairn. Gymnothorax eurostus (Abbott, 1860). BPBM 12247, 220 mm TL, Ducie; BPBM 13236, 2: 441-470 mm TL; BPBM 16492, 190 mm TL, Oeno; BPBM 16563, 4: 245-415 mm TL, Oeno; BPBM 16653, 2: 348-395 mm TL, Pitcairn; BPBM 16738, 492 mm TL, Pitcairn; BPBM 16889, 2: 235-270 mm TL, Pitcairn; BPBM 16968, 103 mm TL, Pitcairn; BPBM 17084, 2: 104-197 mm TL, Henderson. (Fig. 8). Gymnothorax fuscomaculatus (Schultz, 1953). BPBM 16488, 3: 130-160 mm TL, Oeno; BPBM 16624, 4: 131-197 mm TL, Oeno; BPBM 17048, 4: 105- 170 mm TL, Pitcairn. Gymnothorax gracilicauda Jenkins, 1903. BPBM 16462, 426 mm TL, Oeno; BPBM 16558, 125 mm TL, Oeno. (Fig. 9). 7 Gymnothorax javanicus (Bleeker, 1859). BPBM 16626, 1900 mm TL, 24.5 kg (only the head retained), Henderson; one of 1630 mm TL collected at Pitcairn, but not saved; a partly digested Diodon sp., 164 mm SL, was in the stomach. Gymnothorax kidako (Temminck and Schlegel, 1846), 1865. BPBM 16906, 286 mm TL, Pitcairn. (Fig. 10). Identification by Eugenia B. Bohlke. Gymnothorax melatremus Schultz, 1953. BPBM 16620, 9: 73-295 mm TL, Oeno; BPBM 16881, 2: 145-158 mm TL, Pitcairn. Gymnothorax meleagris (Shaw, 1795). BPBM 13237, 845 mm TL, Pitcairn; BPBM 16562, 2: 160-600 mm TL, Oeno; observed at Henderson (Irving et al., 1995). Gymnothorax nasuta de Buen, 1961. BPBM 13265, 673 mm TL, Pitcairn. (Fig. 11). Gymnothorax pindae Smith, 1962. BPBM 16694, 264 mm TL, Pitcairn. Uropterygius fuscoguttatus Schultz, 1953. BPBM 16622, 3: 94-130 mm TL, Oeno; BPBM 16982, 144 mm TL, Pitcairn. Uropterygius inornatus Gosline, 1958. BPBM 16614, 141 mm TL, Oeno. Uropterygius kamar McCosker and Randall, 1977. BPBM 16471, 4: 150-325 mm TL (paratypes), Oeno; BPBM 16606, 297 mm TL (paratype), Oeno ; BPBM 17005, 136 mm TL (paratype), Pitcairn; BPBM 17047, 327 mm TL (paratypes), Pitcairn. (Fig. 12). Uropterygius macrocephalus (Bleeker, 1865). BPBM 16722, 2: 322-433 mm TL, Pitcairn; BPBM 17067, 7: 228-345 mm TL, Oeno; BPBM 17083, 137 mm, Henderson. Identified by John E. McCosker. Uropterygius supraforatus (Regan, 1909). BPBM 16458, 2: 259-378 mm TL, Oeno. (Fig. 13). Uropterygius xanthopterus Bleeker, 1859. BPBM 16460, 4: 171-622 mm TL, Oeno; BPBM 16458, 2: 259-378 mm TL, Oeno; BPBM 16460, 4: 171-622 mm TL, Oeno; BPBM 16470, 275 mm TL, Oeno; BPBM 16619, 9: 180-350 mm TL, Oeno; BPBM 16673, 586 mm TL, Pitcairn; BPBM 16859, 764 mm TL, Pitcairn; BPBM 16885, 348 mm TL, Pitcairn. (Fig. 14). SYNODONTIDAE (LIZARDFISHES) Saurida flamma Waples, 1982. BPBM 13258, 188 mm SL, Pitcairn. Saurida gracilis (Quoy and Gaimard, 1824). BPBM 16481, 78 mm SL, Oeno; BPBM 16792, 3: 145-365 mm SL, Pitcairn; BPBM 17165, 220 mm SL, Ducie; observed at Henderson (Irving et al., 1995). Synodus capricornis Cressey and Randall, 1978. BPBM 16860, 2: 75-103 mm SL (paratypes), Pitcairn. Synodus variegatus (Lacepéde, 1803). BPBM 16768, 183 mm SL, Pitcairn; BPBM 167839, 2: 236-260 mm SL, Pitcairn. ANTENNARIIDAE (FROGFISHES) J Antennarius coccineus (Lesson, 1831). BPBM 11175, 25 mm SL, Oeno; BPBM 28795, 30 mm SL, Oeno. Antennatus tuberosus (Cuvier, 1817). BPBM 7688, 11.5 mm SL, Oeno; BPBM 11176, 3: 30-39 mm SL, Oeno. (Fig. 17). ISONIDAE (SURF FISHES) Iso nesiotes Saeed, Ivantsoff and Crowley, 1993. BPBM 16718, 8: 17-23 mm SL, Pitcairn. OPHIDIIDAE (CUSK EELS) Brotula multibarbata Temminck and Schlegel, 1846. BPBM 16515, 196 mm SL, Oeno Atoll; BPBM 16848, 85 mm TL, Pitcairn. BYTHITIDAE (VIVIPAROUS BROTULAS) Brosmophysiops pautzket Schultz, 1960. BPBM 16548, 2: 26-39 mm SL, Oeno; BPBM 16987, 50 mm SL, Pitcairn. Dinematicthys sp. Ten lots of this species on loan to Yoshihiko Machida. CARAPIDAE (PEARLFISHES) Onuxodon fowleri (Smith, 1955). BPBM 16436, 81 mm TL, Pitcairn. BELONIDAE (NEEDLEFISHES) Ablennes hians (Valenciennes, 1846). Observed by the author off Pitcairn. Platybelone argalus platyura (Bennett, 1832). BPBM 16819, 460 mm FL, Pitcairn; BPBM 17828, 410 mm FL, Henderson; observed at Ducie. Tylosurus crocodilus crocodilus (Peron and Lesueur, 1821). Henderson (Irving et al., 1995). HEMIRAMPHIDAE (HALFBEAKS) Euleptorhamphus viridis (van Hasselt, 1823). BPBM 36422, 440 mm SL, regurgitated by a booby, Henderson. A group of about eight were airborn in front of the ship’s small boat when underway at Pitcairn where the depth was only about 15 m. Hyporhamphus acutus acutus (Giinther, 1861). BPBM 16830, 142 mm SL, Pitcairn; BPBM 17100, 4: 82-143 mm SL, Henderson. EXOCOETIDAE (FLYINGFISHES) Cheilopogon sp. BPBM 16831, 58 mm SL, Pitcairn. Cypselurus pitcairnensis Nichols and Breder, 1935. The holotype (AMNH 12983, 227 mm SL) and four paratypes (AMNH 13293) are in the American Museum of Natural History, New York. Exocoetus obtusirostris Giinther, 1866. BPBM 36424, 55 mm SL, between Henderson and Ducie (found on deck of “Westward”). Exocoetus sp. AMNH 6266, 140 mm SL, from the gullet of a blue-faced booby, Ducie. Reported as Halocypselus evolans (Linnaeus) by Nichols 9 (1923). Specimen borrowed from the American Museum of Natural History by the author and reported as a species of Exocoetus, probably E. volitans (Linnaeus) in Rehder and Randall (1975). The specimen was noted as in very poor condition. HOLOCENTRIDAE (SOLDIERFISHES AND SQUIRRELFISHES) Myrtpristis amaena (Castelnau, 1873). BPBM 12308, 2: 153-201 mm SL, Pitcairn; BPBM 16581, 188 mm SL, Oeno; BPBP 37090, 203 mm SL, Ducie. Myripristis berndti Jordan and Evermann, 1903. BPBM 16580, 2: 56-223 mm SL, Oeno; BPBM 16655, 4: 93-163 mm SL, Pitcairn; BPBM 17140, 231 mm SL, Ducie. Myripristis randalli Greenfield, 1974. BPBM 12116, 136 mm SL (paratype), Pitcairn. Myripristis tiki Greenfield, 1974. BPBM 12115, 14: 127-184 mm SL (paratypes), Pitcairn; BPBM 12307, 153 mm SL, Pitcairn; BPBM 17170, 184 mm SL, Ducie. Neoniphon sammara (Forsskal, 1775). BPBM 17172, 178 mm, Ducie. Plectrypops lima (Valenciennes, 1831). BPBM 16559, 4: 58-88 mm SL, Oeno; BPBM 16643, 2: 72-100 mm SL, Pitcairn; BPBM 16856, 2: 65-100 mm SL, Pitcairn. Sargocentron diadema (Lacepéde, 1801). BPBM 16512, 12: 47-74 mm SL, Oeno; BPBM 16944, 110 mm SL, Pitcairn. Sargocentron ensifer Jordan and Evermann, 1903. BPBM 16787, 194 mm SL, Pitcairn; BPBM 16791, 5: 161-220 mm SL, Pitcairn. Sargocentron hormion Randall, 1998. BPBM 16664, 143 mm SL (holotype), Pitcairn; BPBM 16635, 2: 66-67.5 mm SL (paratypes), Pitcairn; BPBM 16785, 2: 99-145 mm SL (paratypes), Pitcairn. Sargocentron lepros (Allen and Cross, 1983). BPBM 16530, 2: 44-45 mm SL, Oeno; BPBM 16788, 82 mm SL, Pitcairn. Sargocentron megalops Randall, 1998. BPBM 17111, 79 mm SL (holotype), Henderson. Listed as Sargocentron sp. in Irving et al., 1995. Sargocentron punctatissimum (Cuvier, 1829). BPBM 16497, 65 mm SL, Oeno; BPBM 16661, 2: 109-119 mm SL, Pitcairn; BPBM 17053, 5: 93-117 mm SL, Henderson. Sargocentron spiniferum (Forsskal, 1775). BPBM 16618, 346 mm SL (head saved), Oeno; BPBM 17139, 325 mm SL (head saved), Ducie; observed at Henderson (Irving et al., 1995). Sargocentron tiere (Cuvier, 1829). BPBM 6578, 209 mm SL, Oeno; BPBM 16659, 3: 185-210 mm SL, Pitcairn; BPBM 17061, 208 mm SL, 10 Henderson; BPBM 38418, 2: 50-52 mm SL, Oeno; observed at Ducie. AULOSTOMIDAE (TRUMPETFISHES) Aulostomus chinensis (Linnaeus, 1766). BPBM 16752, 470 mm SL, Pitcairn; BPBM 16757, 402 mm SL, Pitcairn; observed at Ducie. FISTULARIIDAE (CORNETFISHES) Fistularia commersoni Rippell, 1838. BPBM 16679, 798 mm SL, Pitcairn; BPBM 16767, 675 mm SL, Pitcairn; BPBM 17118, 752 mm SL, Ducie (misidentified as F. petimba by Rehder and Randall, 1975); observed at Henderson and Oeno. SYNGNATHIDAE (PIPEFISHES AND SEAHORSES) Cosmocampus howensis (Whitley, 1948). BPBM 10856, 2: 80-82 mm SL (paratypes of Syngnathus caldwelli Herald and Randall, 1972), Pitcairn. SCORPAENIDAE (SCORPIONFISHES) Iracundus signifer Jordan and Evermann, 1903. BPBM 11174, 56 mm SL, Oeno; BPBM 11177, 5: 35-39 mm SL, Oeno; BPBM 11208, 2: 36-37 mm SL, Pitcairn; BPBM 11217, 35 mm SL, Pitcairn; BPBM 11230, 33 mm SL, SL, Pitcairn; BPBM 16888, 34-37 mm SL, Pitcairn. (Fig. 18). Parascorpaena mcadamsi (Fowler, 1938). BPBM 11195, 5: 32-43 mm SL, Pitcairn; BPBM 11234, 5: 25.7-48.9 mm SL, Ducie. (Fig. 19) Pontinus sp. BPBM 16439, 53 mm SL, Pitcairn; BPBM._ 16448, 37 mm SL, Pitcairn, both from dredging in 92-128 m, 19-20 October, 1967; identified by William N. Eschmeyer. Pterois antennata (Bloch, 1787). BPBM 13261, 152 mm SL, Pitcairn; BPBM 16627, 114 mm SL, Oeno; BPBM 16636, 100 mm SL, Pitcairn; observed at Henderson (Irving et al., 1995). Pterois volitans (Linnaeus, 1758). BPBM 16629, 193 mm SL, Oeno; observed at Henderson (Irving et al., 1995). Scorpaenodes hirsutus (Smith, 1957). BPBM 11178, 2: 13-45 mm SL, Oeno; BPBM 11204, 2: 40-43 mm SL, Pitcairn. (Fig. 20). Scorpaenopsis fowleri (Pietschmann, 1934). BPBM 11181, 7: 23-30 mm SL, Oeno. This small species will probably be reclassified in a new genus. Scorpaenopsis sp. BPBM 16770, 2: 175-195 mm SL, Pitcairn. These specimens represent a new species; currently on loan to William N. Eschmeyer. Sebastapistes galatacma Jenkins, 1903. BPBM 16447, 18 mm SL, Pitcairn; BPBM 16605, 38 mm SL, Oeno; BPBM 16945, 19 mm SL, Pitcairn. Sebastapistes mauritiana (Cuvier, 1829). BPBM 11198, 3: 41-51 mm SL, Pitcairn; BPBM 11213, 3: 20-34 mm, Pitcairn; BPBM 11236, 3: 20-33 mm SL, Henderson; BPBM 17077, 17 mm SL, Henderson. (Fig. 21). Sebastapistes tinkhami (Fowler, 1946). BPBM 11179, 4: 39-67 mm SL, Oeno; BPBM 11199, 9: 18-40 mm SL, Pitcairn; BPBM 11206, 3: 31-43 mm SL, ERRATUM NO. 461 REPORT ON FISH COLLECTIONS FROM THE PITCAIRN ISLANDS BY JOHN E. RANDALL Page 7 Uropterygius xanthopterus Bleeker, 1859 should be recorded as Uropterygius alboguttatus Smith, 1961 1] Pitcairn; BPBM 11211, 5: 37-69 mm SL, Pitcairn; BPBM 11216, 6: 30-43 mm SL, Pitcairn; BPBM 11235, 10: 29-49 mm SL, Henderson; BPBM 16491, 2: 35-42 mm SL, Oeno. PLATYCEPHALIDAE (FLATHEADS) Eurycephalus otaitensis (Cuvier, 1829): BPBM 16696, 3: 62-173 mm SL, Pitcairn; BPBM 16782, 3: 103-137 mm SL, Pitcairn; BPBM 17034, 72 mm SL, Pitcairn. CARACANTHIDAE (ORBICULAR VELVETFISHES) Caracanthus maculatus (Gray, 1831). BPBM 16539, 49 mm SL, Oeno; BPBM 16630, 3: 40-48 mm SL, Oeno; BPBM 16929, 4: 31-44 mm SL, Pitcairn. (Fig. 22). Caracanthus untpinna (Gray, 1831). BPBM 16505, 16: 13-26 mm SL, Oeno; BPBM 17098, 12: 17-29 mm SL, Henderson. DACTYLOPTERIDAE (HELMET GURNARDS) Dactyloptena orientalis (Cuvier, 1829). BPBM 16833, 235 mm SL, Pitcairn. SERRANIDAE (GROUPERS AND SEABASSES) Cephalopholis argus Bloch and Schneider, 1801. BPBM 16577, 245 mm SL, Oeno; BPBM 16666, 173 mm SL, Pitcairn; BPBM 17162, 223 mm SL, Ducie; observed at Henderson (Irving et al., 1995). Cephalopholis spiloparaea (Valenciennes, 1828). BPBM 13260, 5: 108-164 mm SL, Pitcairn; BPBM 16790, 3: 47-145 mm SL, Pitcairn. (Fig 23). Cephalopholis urodeta (Forster in Bloch and Schneider, 1801). BPBM 16566, 3: 98-114 mm SL, Oeno; BPBM 16654, 192 mm SL, Pitcairn; observed at Henderson (Irving et al., 1995). Epinephelus fasciatus (Forsskal, 1775). BPBM 16610, 65 mm SL, Oeno; BPBM 17045, 69 mm SL, Pitcairn; BPBM 17126, 219 mm SL, Ducie. Eptnephelus hexagonatus (Forster in Bloch and Schneider, 1801). BPBM 13240, 2: 123-211 mm SL, Pitcairn; BPBM 17055, 108 mm SL, Henderson. Epinephelus lanceolatus (Bloch, 1790). BPBM 33914, 1450 mm SL, 148 kg (saved head), Henderson; collected by the author with powerhead and spear. (Fig. 24). Pitcairn Miscellany reported two large E. lanceolatus landed at Pitcairn, one of over 182 cm total length speared by Steve Christian in January 1991 and one in January 1993. Epinephelus merra Bloch, 1793. Observed at Henderson (Irving et al., 1995). Epinephelus socialis (Giinther, 1873). BPBM 16809, 2: 125-197 mm SL, Pitcairn; BPBM 17099, 73 mm SL, Henderson; BPBM 17150, 242 mm SL, Ducie. Epinephelus tauvina (Forsskal, 1775). BPBM 12256, 84 mm SL, Ducie; BPBM 12 16579, 278 mm SL, Oeno; BPBM 16698, 2: 174-265 mm SL, Pitcairn; BPBM 16724, 3: 97-113 mm SL, Pitcairn; BPBM 16807, 3: 117-151 mm SL, Pitcairn; BPBM 17054, 86 mm SL, Henderson. Epinephelus tuamotuensis Fourmanoir, 1971. BPBM 17110, 398 mm SL (hook and line from 120 m), Henderson. Liopropoma pallidum (Fowler, 1838). BPBM 16543, 2: 47-63.5 mm SL, Oeno. Plectranthias fourmanoiri Randall, 1980. BPBM 15068, 36.6 mm SL (paratype), Pitcairn; Plectranthias nanus Randall, 1980. BPBM 16542, 3: 22-37.2 mm SL (paratypes), Oeno; BPBM 16744, 2: 25-27.5 mm SL (paratypes), Pitcairn; BPBM 16847, 30.8 mm SL (paratype), Pitcairn; BPBM 16891, 2: 31.1-34.8 mm SL (paratypes), Pitcairn; BPBM 17027, 35.6 mm SL (paratype), Pitcairn; BPBM 17033, 31 mm SL, Pitcairn. Plectranthias winniensis (Tyler, 1966). BPBM 16743, 31 mm SL, Pitcairn; BPBM 16901, 33 mm SL, Pitcairn; also paratypes from Ducie were deposited in the California Academy of Sciences and the U.S. National Museum of Natural History. Plectropomus laevis (Lacepéde, 1801). Observed at Oeno. Pseudanthias mooreanus (Herre, 1935). BPBM 16466, 72 mm SL, Oeno; BPBM 16478, 2: 68-69 mm SL, Oeno; BPBM 16613, 10: 34-64 mm SL, Oeno; BPBM 16854, 3: 48-52 mm SL, Pitcairn. Pseudanthias ventralis (Randall, 1979). BPBM 16741, 19 mm SL, Pitcairn; BPBM 16862, 2: 32-35 mm SL, Pitcairn Island; BPBM 16883, 51.5 mm SL (holotype), Pitcairn; BPBM 16892, 6: 23-40 mm SL, Pitcairn; observed at Henderson (Irving et al., 1995); abundant in 37-46 m at Ducie. Pseudogramma australis Randall and Baldwin, 1997. BPBM_ 16443, 41 mm SL, Pitcairn; BPBM 16444, 27 mm SL (paratype), Pitcairn; BPBM 16893, 33.1 mm SL (paratype), Pitcairn. Pseudogramma polyacanthum (Bleeker, 1856). BPBM 12272, 4: 37-56 mm SL, Ducie; BPBM 16494, 14: 29-59 mm SL, Oeno; BPBM 16552, 11: 19.5-60.5 mm SL, Oeno; BPBM 16973, 9: 35-56 mm SL, Pitcairn. Variola louti (Forsskal, 1775). BPBM 16574, 536 mm SL (saved head), Oeno; BPBM 16745, 118 mm SL, Pitcairn; BPBM 16756, 234 mm, Pitcairn; BPBM 17051, 147 mm SL, Henderson; BPBM 17131, 2: 193-196 mm SL, Ducie. KUHLIDAE (FLAGTAILS) Kuhlia marginata (Cuvier, 1829). BPBM 12250, 2: 160-162 mm, Ducie; BPBM 16737, 8: 108-166 mm SL, Pitcairn; BBPM 17063, 3: 25-158 mm SL, Henderson; BPBM 17125, 2: 208-215 mm SL, Ducie. The identification of this species as K. marginata is provisional; a revision of the genus is needed. PSEUDOCHROMIDAE (DOTTYBACKS) Pseudoplesiops revellet Schultz, 1953. BPBM 16484, 3: 35-36 mm SL, Oeno. PRIACANTHIDAE (BIGEYES) Heteropriacanthus cruentatus (Lacepéde, 1801). BPBM 16665, 236 mm SL, Pitcairn; BPBM 17169, 260 mm SL, Ducie. CIRRHITIDAE (HAWKFISHES) Amblycirrhitus bimacula (Jenkins, 1903). BPBM 16523, 2: 50-63 mm SL, Oeno; BPBM 16633, 3: 30-57 mm SL, Pitcairn; BPBM 17030, 48 mm SL, Pitcairn. Amblycirrhitus wilhelmi (Lavenberg and Ydafiez, 1972). BPBM 16504, 5: 40-64 mm SL, Oeno; BPBM 16637, 68 mm SL, Pitcairn; BPBM 16638, 13: 41- 120 mm SL, Pitcairn; BPBM 16740, 38 mm SL, Pitcairn; BPBM 16609, 98 mm SL, Pitcairn; BPBM 16956, 53 mm SL, Pitcairn; BPBM 17068, 9: 45- 111 mm SL, Henderson. Placement of this species in the genus Amblycirrhitus is provisional; it may require a new genus. Cirrhitops hubbardi (Schultz, 1943). BPBM 16642, 78 mm SL, Pitcairn; BPBM 17105, 75 mm SL, Henderson; observed at Oeno. Cirrhitus pinnulatus (Forster in Bloch and Schneider, 1801). BPBM 16713, 155 mm SL, Pitcairn; observed at Henderson. Neocirrhites armatus Castelnau, 1873. BPBM 16446, 20 mm SL, Pitcairn; BPBM 16502, 6: 49-64 mm SL, Oeno; BPBM 16644, 4: 53-80 mm SL, Pitcairn. Paracirrhites arcatus (Cuvier, 1829). BPBM 16551, 16: 34-105 mm SL, Oeno; BPBM 16761, 108 mm SL, Pitcairn; common at Henderson (Irving et al., 1995). Paracirrhites forsteri (Bloch and Schneider, 1801). BPBM 13242, 142 mm SL, Pitcairn; BPBM 16651, 167 mm SL, Pitcairn; BPBM 16663, 135 mm SL, Pitcairn; BPBM 17153, 142 mm SL, Ducie; common at Henderson (Irving et al., 1995). Paracirrhites hemistictus (Gunther, 1874). BPBM 16465, 215 mm SL, Oeno; BPBM 16634, 62 mm SL, Pitcairn; BPBM 16714, 206 mm SL, Pitcairn; BPBM 17137, 229 mm SL, Ducie; common at Henderson (Irving et al., 1995). Paracirrhites nisus Randall, 1963. BPBM 16521, 93 mm SL, Oeno. APOGONIDAE (CARDINALFISHES) Apogon angustatus (Smith and Radcliffe, 1911). BPBM 16469, 57 SL, Oeno; BPBM 16522, 8: 48-87 mm SL, Oeno; BPBM 16919, 2: 70-74 mm SL, Pitcairn; BPBM 17037, 88 mm SL, Pitcairn. Apogon caudicinctus Randall and Smith, 1988. BPBM 16799, 2: 53.9-64.3 mm SL (paratypes), Pitcairn. Apogon kallopterus Bleeker, 1856. BPBM 16490, 2: 45-67 mm SL, Oeno; BPBM 16652, 102 mm SL, Pitcairn. Apogon taeniophorus Regan, 1908. BPBM 12258, 6: 14-70 mm SL, Ducie; BPBM 16479, 7: 42-62 mm SL, Oeno; BPBM 16726, 6: 69-79.4 mm SL, Pitcairn; BPBM 16806, 6: 70-101 mm SL, Pitcairn. Apogon taentopterus Bennett, 1836. BPBM 16493, 3: 55-65 mm SL, Oeno; BPBM 16686, 5: 100-126 mm SL, Pitcairn; BPBM 16898, 107 mm SL, Pitcairn; BPBM 17039, 123 mm SL, Pitcairn. Apogon sp. Seven lots of a small transparent red species of the Apogon coccineus complex are under study by David W. Greenfield. These specimens have 14 pectoral rays and usually 3 + 12 gill rakers. Apogon sp. BPBM 16895, 36 mm SL, from 55 m at Pitcairn. An undescribed species that has been sent on loan to Thomas H. Fraser. It is characterized by VII-I,9 dorsal rays II,8 anal rays, 14 pectoral rays, 6 + 17 gill rakers, a straight dorsal head profile, smooth preopercular ridge, and all pale coloration except for a small dusky spot at midbase of caudal fin. Cercamia cladara Randall and Smith, 1988. BPBM 16472, 3: 35-36 mm SL, Oeno; BPBM 16954, 2: 35 mm SL, Pitcairn. Cheilodipterus macrodon (Lacepéde, 1801). BPBM 16783, 136.5 mm SL, Pitcairn; BPBM 16784, 166.2 mm SL, Pitcairn. Chetlodipterus quinquelineatus Cuvier, 1828. BPBM 17173, 54 mm SL, Ducie. Gymnapogon vanderbilti (Fowler, 1938). BPBM 16855, 24 mm SL, Pitcairn. Gymnapogon sp. BPBM 16978, 23 mm SL, Pitcairn. An undescribed species under study by Jeng-Ping Chen, who provided the identification of the other Pitcairn species of the genus. Pseudamiops gracilicauda (Lachner in Schultz, 1953). BPBM 16485, 19 mm SL, Oeno. ECHENEIDAE (REMORAS) Phtheirichthys lineatus (Menzies, 1791). BPBM 16953, 85 mm SL, Pitcairn. Found in the bottom of the dive boat after spearing a specimen of Ditodon hystrix, 275 mm SL, and several spiny lobsters. Remora remora (Linneaus, 1758). AMS 1.4252, 73 mm SL, obtained by the Australian Museum on exchange from R.E. Vaughan, June, 1899. CARANGIDAE (JACKS) Carangoides ferdau (Forsskal, 1775). BPBM 16677, 353 mm FL, Pitcairn; 15 observed at Henderson and Ducie. Carangoides orthogrammus (Jordan and Gilbert, 1881). BPBM 16585, 518 mm FL (saved head), Pitcairn. Observed at Henderson (Irving et al., 1995). Caranx ignobilis (Forsskal, 1775). A specimen 1190 mm FL, 38 kg, collected at Henderson but not retained; also observed at Oeno and Ducie. Caranx lugubris Poey, 1860. Nine caught at Oeno to 675 mm FL, retained the smallest, BPBM 16589, 360 mm FL; BPBM 16671, 422 mm FL (head saved), Pitcairn; BPBM 16777, 290 mm FL, Pitcairn; BPBM 17161, 252 mm FL, Ducie. Caranx melampygus (Cuvier, 1833). BPBM 16573, 560 mm FL (saved head), Oeno; BPBM 17075, 315 mm FL, Henderson; observed at Ducie and Pitcairn (not common at Pitcairn). Decapterus sp. Observed at Oeno. Pseudocaranx dentex (Bloch and Schneider, 1801). BPBM 13267, 302 mm FL, Pitcairn; BPBM 17109, 133 mm FL, Henderson; observed at Oeno. Seriola lalandi Valenciennes, 1833. BPBM 16965, 915 mm SL (saved head and caudal fin), Pitcairn; observed at Henderson (Irving et al., 1995), and illustrated at Ducie with an underwater photograph (Rehder and Randall, 1975: fig. 28). (Fig. 25). Seriola rivoliana Valenciennes, 1833. BPBM 17114, 500 mm FL, Ducie. Uraspis sp. BPBM 17295, 337 mm FL, Pitcairn. This specimen was loaned to Frank Williams and later transferred to Frederick H. Berry. It now appears to be lost. CORYPHAENIDAE (DOLPHINS) Coryphaena hippurus Linnaeus, 1758. BPBM 36429, 93 mm SL (dropped on land by a fairy tern), Henderson. LUTJANIDAE (SNAPPERS) Aphareus furca (Lacepéde, 1802). BPBM 16817, 352 mm SL (saved head), Pitcairn; observed at Henderson and Oeno. Lutjanus bohar (Forsskal, 1775). BPBM 16603, 360 mm SL, Ducie; five caught at Oeno, the largest 720 mm SL, 16.3 kg, none retained; observed at Pitcairn and Henderson. Lutjanus kasmira (Forsskal, 1775). BPBM 16603, 255 mm SL, Oeno; BPBM 16772, 260 mm SL, Pitcairn. Lutjanus monostigma (Cuvier, 1828). One of about 350 mm TL speared at Pitcairn, but escaped. Paracaesio sordidus Abe and Shinohara, 1962. BPBM 16774, 2: 233-240 mm SL, Pitcairn. (Fig. 26). CAESIONIDAE (FUSILIERS) Pterocaesio tile (Cuvier, 1830). Observed at Pitcairn; also at Henderson 16 (Irving et al., 1995). LETHRINIDAE (EMPERORS) Gnathodentex aureolineatus (Lacepéde, 1802). BPBM 13257, 224 mm SL, Pitcairn; BPBM 17119, 202 mm SL, Ducie; observed at Henderson (Irving et al., 1995). Lethrinus olivaceus Valenciennes, 1830. One observed at Henderson (identified as L. elongatus by Irving et al., 1995). Monotaxis grandoculis (Forsskal, 1775). BPBM 16463, 225 mm SL, Oeno; BPBM 16519, 80 mm SL, Oeno; two of 385 and 440 mm SL speared at Pitcairn, the largest 2.4 kg, neither retained; observed at Ducie and Henderson. MUGILIDAE (MULLETS) Neomyxus leuciscus (Giinther, 1871). BPBM 16464, 12: 90-242 mm SL, Oeno; BPBM 16728, 15: 85-140 mm SL, Pitcairn; observed in shoals on the reef platform at Henderson (Irving et al., 1995); also observed at Ducie. POLYNEMIDAE (THREADFINS) Polydactylus sexfilis (Valenciennes, 1831). BPBM 16828, 2: 240-268 mm SL, Pitcairn. MULLIDAE (GOATFISHES) Mulloidichthys flavolineatus (Lacepéde, 1801). BPBM 16657, 214 mm SL, Pitcairn; BPBM 17151, 298 mm SL, Ducie; reported from Oeno by Nichols (1923) as Mulloides samoensis; observed at Henderson. Mulloidichthys vanicolensis (Valenciennes, 1831). BPBM 13266, 172 mm SL, Pitcairn; BPBM 16576, 2: 236-237 mm SL, Oeno; observed at Henderson (Irving et al., 1995). Parupeneus bifasciatus (Lacepeéde, 1801). BPBM 16586, 2: 203-236 mm SL, Oeno; BPBM 16747, 129 mm SL, Pitcairn; BPBM 16938, 154 mm SL, Pitcairn; BPBM 17058, 2: 114-117 mm SL, Henderson; BPBM 17148, 214 mm SL, Ducie. Parupeneus ciliatus (Lacepéde, 1802). BPBM 16753, 312 mm SL, Pitcairn. Parupeneus cyclostomus (Lacepéde, 1801). BPBM 13238, 2: 145-290 mm SL, Pitcairn; BPBM 16587, 226 mm SL, Oeno; BPBM 17132, 310 mm SL, Ducie; observed at Henderson (Irving et al., 1995). Parupeneus multifasciatus (Quoy and Gaimard, 1825). BPBM 13253, 260 mm SL, Pitcairn; BPBM 13263, 150 mm SL, Pitcairn; BPBM 16561, 3: 90-103 mm SL, Oeno; BPBM 16639, 123 mm SL, Pitcairn; BPBM 16716, 230 mm SL, Pitcairn; BPBM 17130, 244 mm SL, Ducie; observed at Henderson (as P. trifasciatus) by Irving et al. (1995). Parupeneus pleurostigma (Bennett, 1831). BPBM 13264, 6: 117-225 mm SL, 17 Pitcairn; BPBM 16924, 65 mm SL, Pitcairn; observed at Ducie. PEMPHERIDAE (SWEEPERS) Pempheris otaitensis (Cuvier, 1831). BPBM 13252, 2: 165-181 mm SL, Pitcairn; BPBM 16804, 2: 90-120 mm SL, Pitcairn; BPBM 17074, 133 mm SL, Henderson. KYPHOSIDAE (SEA CHUBS) Kyphosus bigibbus Lacepeéde, 1802. BPBM 16712, 204 mm SL, Pitcairn; BPBM 16736, 170 mm SL, Pitcairn; BPBM 16835, 2: 222-390 mm SL, Pitcairn. Common at all the islands; reported from Ducie as K. fuscus by Rehder and Randall (1975). EPHIPPIDAE (SPADEFISHES) Platax sp. One sight record by Irving et al., 1995 at Henderson; indicated as probably P. orbicularis (Forsskal, 1775). CHAETODONTIDAE (BUTTERFLYLFISHES) Chaetodon auriga Forsskal, 1775. BPBM 17127, 169 mm SL, Ducie; observed at the other three islands. Chaetodon flavirostris Gunther, 1874. BPBM 16592, 2: 153-154 mm SL, Oeno; BPBM 17107, 162 mm SL, Henderson; BPBM 17120, 57 mm SL, Ducie. Chaetodon lunula (Lacepéde, 1802). BPBM 16590, 2: 152-160 mm SL, Oeno; BPBM 16676, 148 mm SL, Pitcairn; BPBM 16735, 83 mm SL, Pitcairn; observed at Henderson. Chaetodon mertensiit Cuvier, 1831. BPBM 13243, 4: 43-118 mm SL, Pitcairn; BPBM 16560, 93 mm SL, Oeno; BPBM 17176, 97 mm SL, Ducie; observed at Henderson (Irving et al., 1995). Chaetodon ornatissimus Cuvier, 1831. BPBM 13248, 79 mm SL, Pitcairn; BPBM 16508, 138 mm SL, Oeno; BPBM 17154, 2: 140-142 mm SL, Ducie; underwater photo taken at Henderson. Chaetodon pelewensis Kner, 1868. BPBM 13245, 3: 70-85 mm SL, Pitcairn; BPBM 16555, 4: 61-75 mm SL, Oeno; BPBM 16937, 74 mm SL, Pitcairn; noted as the most common butterflyfish at Henderson by Irving et al. (1995); also common at Ducie. Chaetodon quadrimaculatus Gray, 1831. BPBM 12238, 120 mm SL, Ducie; BPBM 12277, 120 mm SL, Henderson; BPBM 16567, 2: 111-114 mm SL, Oeno; BPBM 16762, 111 mm SL, Pitcairn. Chaetodon reticulatus Cuvier, 1831. BPBM 12243, 119 mm SL, Ducie; BPBM 16765, 113 mm SL, Pitcairn; observed at Henderson. Chaetodon smitht Randall, 1975a. BPBM 13216, 128.7 mm SL (paratype), Pitcairn; BPBM 13217, 123.2 mm SL (paratype), Pitcairn; BPBM 13218, 126.1 mm SL (paratype), Pitcairn; BPBM 13220, 129.8 mm SL (holotype), Pitcairn; BPBM 13221, 142 mm SL (paratype), Pitcairn. Chaetodon ulietensis Cuvier, 1831. BPBM 17175, 2: 136-145 mm SL, Ducie; the most common butterflyfish at the atoll. Chaetodon unimaculatus Bloch, 1787. BPBM 16758, 164 mm SL, Pitcairn; BPBM 17117, 101 mm SL, Ducie; observed at Henderson (Irving et al., 1995): Forcipiger flavissimus Jordan and McGregor, 1898. BPBM 16649, 2: 118-123 mm SL, Pitcairn; BPBM 17163, 165 mm SL, Ducie. Forcipiger longirostris (Broussonet, 1782). BPBM 16700, 196 mm SL, Pitcairn; BPBM 16787, 2: 159-171 mm SL, Pitcairn. Both species of the genus observed at Henderson (Irving et al., 1995). Hemitaurichthys multispinosus Randall, 1975a. BPBM 13222, 2: 136.2-159 mm SL (paratypes), Pitcairn; BPBM 13225, 144.1 mm SL (holotype), Pitcairn; BPBM 13327, 2: 136-155 mm SL (paratypes), Pitcairn; BPBM 20802, 2: 146-147 mm SL, Pitcairn. (Fig. 27). Hemitaurichthys polylepis (Bleeker, 1857). BPBM 16701, 125 mm SL, Pitcairn. Heniochus chrysostomus Cuvier, 1831. BPBM 16591, 2: 124-129 mm SL, Oeno; BPBM 16814, 138 mm SL, Pitcairn. Heniochus monoceros Cuvier, 1831. BPBM 16601, 184 mm SL, Oeno; observed at Henderson (Irving et al., 1995). POMACANTHIDAE (ANGELFISHES) Centropyge flavissima (Cuvier, 1831). BPBM 12240, 83 mm SL, Ducie; BPBM 13241, 5: 43-60 mm SL, Pitcairn; BPBM 16940, 2: 44-71 mm SL, Pitcairn; BPBM 30201, 2: 39-62 mm SL, Oeno; observed at Henderson (Irving et al., 1995). Centropyge heraldi Woods and Schultz, 1953. BPBM 13262, 2: 92-95 mm SL, Pitcairn; BPBM 16518, 3: 40-67 mm SL, Oeno. Centropyge hotumatua Randall and Caldwell, 1973. BPBM 12275, 3: 32-48 mm SL (paratypes), Ducie; BPBM 13312, 5: 24.9-61.8 mm SL (paratypes), Pitcairn; BPBM 13314, 2: 27.8-36.8 mm SL (paratypes), Oeno; BPBM 13315, 3: 24.5-46.2 mm SL (paratypes), Pitcairn; BPBM 13325, 2: 22.6-27.5 mm SL (paratypes; mistakenly listed as BPBM 13326 by Randall and Caldwell, 1973), Pitcairn; observed at Henderson. (Fig. 28). Centropyge loricula (Gunther, 1874). BPBM 16520, 16: 28-52 mm SL, Oeno; BPBM 16843, 2: 44-45 mm SL, Pitcairn; BPBM 16878, 53 mm SL, Pitcairn; BPBM 17171, 37 mm SL, Ducie; observed at Henderson. Genicanthus spinus Randall, 1975b. BPBM 16450, 203 mm SL (holotype), 19 Pitcairn; BPBM 16452, 2: 182.1-199 mm SL (paratypes), Pitcairn; BPBM 16754, 156.2 mm SL (paratype), Pitcairn; one was speared in 55 m off Ducie, but escaped. Genicanthus watanabei (Yasuda and Tominaga, 1970). BPBM 16449, 2: 109- 137 mm SL, Pitcairn; BPBM 16755, 107 mm SL, Pitcairn; observed at Henderson (Irving et al., 1995). Pomacanthus imperator (Bloch, 1787). BPBM 16803, 147 mm SL, Pitcairn; BPBM 17294, 255 mm SL, Pitcairn; underwater photo taken at Henderson. POMACENTRIDAE (DAMSELFISHES) Abudefduf sordidus Forsskal, 1775. BPBM 17022, 3: 90-110 mm SL, Pitcairn; BPBM 27990, 125 mm SL, Ducie. Chromis acares Randall and Swerdloff, 1973. BPBM 16540, 11: 34-44 mm SL, Oeno. Chromis agilis Smith, 1960. BPBM 16525, 6: 37-66 mm SL, Oeno; BPBM 16864, 36 mm SL, Pitcairn; BPBM 17043, 2: 36-70 mm SL, Pitcairn; observed at Henderson and Ducie. Chromis bami Randall and McCosker, 1992. BPBM 16524, 5: 37.6-55 mm SL (paratypes), Oeno; BPBM 16687, 9: 28.3-55.8 mm SL (paratypes), Pitcairn; observed at Henderson (Irving et al., 1995). Chromis pamae Randall and McCosker, 1992. BPBM 16457, 10: 41-86.9 mm SL (paratypes), Oeno; BPBM 16905, 103.8 mm SL (holotype), Pitcairn; BPBM 16922, 5: 48.6-101.7 mm SL (paratypes), Pitcairn; observed at Henderson (Irving et al., 1995). Chromis vanderbilti (Fowler, 1941). BPBM 16914, 3: 40-44 mm SL, Pitcairn; BPBM 16960, 3: 22-34 mm SL, Pitcairn; observed at Ducie. Chromis xanthura (Bleeker, 1854). BPBM 16628, 115 mm SL, Oeno; BPBM 16824, 2: 115-119 mm SL, Pitcairn. Chrysiptera galba (Allen and Randall, 1974). BPBM 11132, 3: 48.5-57.2 mm SL (paratypes), Ducie; BPBM 11173, 50 mm SL, Oeno; BPBM 11183, 10: 34.8-59.3 mm SL, Oeno; BPBM 11184, 4: 37-56 mm SL, Oeno; BPBM 11188, 11: 18-60 mm SL, Oeno; BPBM 11197, 5: 38.2-44.3 mm SL (paratypes), Pitcairn; BPBM 11210, 6: 27-56 mm SL, Pitcairn; BBPM 11214, 5: 33-40 mm SL, Pitcairn; BPBM 11231, 4: 41-57 mm SL, Ducie; observed at Henderson. Dascyllus flavicaudus H. Randall and Allen, 1977. BPBM 16568, 6: 52.7-77 mm SL (paratypes), Oeno; BPBM 16926, 66.1 mm SL (paratype; mistakenly numbered as BPBM 19626 in the original description), Pitcairn; BPBM 16646, 5: 48.7-56.4 mm SL (paratypes), Pitcairn; observed at Henderson (Irving et al., 1995). 20 Dascyllus reticulatus (Richardson, 1846). BPBM 19940, 36 mm SL, Pitcairn. Dascyllus trimaculatus (Riippell, 1829). BPBM 16812, 106 mm SL, Pitcairn. Plectroglyphidodon imparipennis (Vaillant and Sauvage, 1875). BPBM 12252, 8: 33-41 mm, Ducie; BPBM 16867, 3: 18-24 mm SL, Pitcairn; BPBM 16995, 32: 21-27 mm SL, Pitcairn; BPBM 17085, 6: 18-46 mm SL, Henderson. Plectroglyphidodon johnstonianus Fowler and Ball, 1924. BPBM 16513, 4: 50- 75mm SL, Oeno; BPBM 16689, 51 mm SL, Pitcairn; BPBM 16918, 2: 29- 35 mm SL, Pitcairn; BPBM 16934, 66 mm SL, Pitcairn; BPBM 17049, 48 mm SL, Henderson. Plectroglyphidodon leucozona (Bleeker, 1859). BPBM 16727, 15: 58-112 mm SL, Pitcairn; BPBM 17001, 4: 48-67 mm SL, Pitcairn; BPBM 17015, 60 mm SL, Pitcairn; BPBM 17052, 86 mm SL, Henderson. (Fig. 29). Plectroglyphidodon phoenixensis (Schultz, 1943). BPBM 17060. 4: 35-50 mm SL, Henderson; observed at Oeno. Pomachromis fuscidorsalis Allen and Randall, 1974. BPBM 14247, 49 mm SL (paratype), Oeno; BPBM 16528, 45 mm SL, Oeno; BPBM 16692, 3: 35- 40 mm SL, Pitcairn; BPBM 17166, 46 mm, Ducie; observed at Henderson (Irving et al., 1995). Stegastes emeryi (Allen and Randall, 1974). BPBM 14245, 44.2 mm SL (holotype), Pitcairn; BPBM 14246, 3: 48.2-55.9 mm SL (paratypes), Oeno; AMS 1.17346-001, 2: 66.3-67.5 mm, Ducie (paratypes); BPBM 16483, 67 mm SL, Oeno; BPBM 16535, 2: 50-61 mm SL, Oeno; BPBM 16943, 34 mm SL, Pitcairn; BPBM 17050, 2: 56-59 mm SL, Henderson. Stegastes fasciolatus (Ogilby, 1889). BPBM 12248, 2: 60-71 mm SL, Ducie; BPBM 16556, 2: 66-81 mm SL, Oeno; BPBM 16609, 47 mm SL, Oeno; BPBM 16693, 2: 66-78 mm SL, Pitcairn; BPBM 16925, 3: 24-27 mm SL, Pitcairn; BPBM 16936, 2: 22-95 mm SL, Pitcairn; BPBM 16964, 6: 19-59 mm SL, Pitcairn; BPBM 17038, 118 mm SL, Pitcairn; BPBM 17082, 2: 22-23 mm SL, Henderson; BPBM 38414, 79 mm SL, Oeno. LABRIDAE (WRASSES) Anampses caeruleopunctatus Riuppell, 1828. BPBM 12048, 2: 238-260 mm SL, Pitcairn; BPBM 12049, 216 mm SL, Pitcairn; BPBM 12245, 154 mm SL, Ducie; BPBM 12278, 177 mm SL, Henderson; BPBM 16746, 216 mm SL, Pitcairn; BPBM 16818, 103 mm SL, Henderson. Anampses femininus Randall, 1972. BPBM 11602, 2: 37-88 mm SL (paratypes), Pitcairn; BPBM 16480, 55 mm SL, Oeno; BPBM 16647, 150 mm SL, Pitcairn; BPBM 16951, 36 mm SL, Pitcairn. Anampses twistit Bleeker, 1856. BPBM 16670, 94 mm SL, Pitcairn. 21 Bodianus anthioides (Bennett, 1832). Observed at Henderson (Irving et al., 1995). Bodianus axillaris (Bennett, 1831). BPBM 13250, 49 mm SL, Pitcairn; BPBM 13259, 147 mm SL, Pitcairn; BPBM 16516, 150 mm SL, Oeno; BPBM 16685, 140 mm SL, Pitcairn; observed at Henderson. Bodianus bilunulatus (Lacepéde, 1801). BPBM 16588, 265 mm SL, Henderson; BPBM 17133, 4: 220-295 mm SL, Ducie. Cheilinus undulatus Rippell, 1835. One observed in 30 m at Henderson (Irving et al., 1995). Cheilio inermis (Forsskal, 1775). BPBM 16775, 260 mm SL, Pitcairn. Cirrhliabrus scottor'um Randall and Pyle, 1989. BPBM 16474, 59.8 mm SL (paratype), Oeno; BPBM 16739, 85 mm SL, Pitcairn; BPBM 16904, 2: 60.3-62.7 mm SL (paratypes), Pitcairn. Coris aygula Lacepéde, 1801. BPBM 16572, 3: 112-257 mm SL, Oeno; BPBM 16750, 475 mm SL, Pitcairn; BPBM 16810, 3: 95-230 mm SL, Pitcairn; BPBM 17159, 312 mm SL, Ducie; observed at Henderson. Coris sp. 11 lots from Pitcairn and Oeno; a new species to be described by the author. Gomphosus varius Lacepéde, 1801. BPBM 16501, 34 mm SL, Oeno; BPBM 16723, 140 mm SL, Pitcairn; observed at Ducie and Henderson. Halichoeres margaritaceus (Valenciennes, 1839). BPBM 16719, 103 mm SL, Pitcairn; observed at Oeno. Halichoeres marginatus Rippell, 1835. BPBM 16496, 55 mm SL, Oeno. Halichoeres melasmapomus Randall, 1980. BPBM 16702, 2: 107.7-118.2 mm SL (paratypes), Pitcairn; BPBM 16899, 68 mm SL (holotype), Pitcairn. Halichoeres trimaculatus (Quoy and Gaimard, 1834). One observed in the Ducie lagoon. Hemigymnus fasciatus (Bloch, 1792). BPBM 16715, 213 mm SL, Pitcairn; observed at Henderson and Ducie. Hologymnosus annulatus (Lacepéde, 1801). BPBM 16766, 2: 259-315 mm SL, Pitcairn; BPBM 16826, 295 mm SL, Pitcairn. Labroides bicolor Fowler and Bean, 1928. BPBM 13247, 95 mm SL, Pitcairn; observed at Oeno. Labroides dimidiatus (Valenciennes, 1839). BPBM 16546, 2: 56-83 mm SL, Oeno; BPBM 16683, 77 mm SL, Pitcairn; BPBM 16962, 14 mm SL, Pitcairn; BPBM 17040, 105 mm SL, Pitcairn; BPBM 17103, 76 mm SL, Henderson; BPBM 17174, 44 mm SL, Ducie. Labroides rubrolabiatus Randall, 1958. BPBM 13246, 67 mm SL, Pitcairn; BPBM 16545, 2: 51-66 mm SL, Oeno; BPBM 16687, 4: 42-70 mm SL, 22 Pitcairn; BPBM 17168, 2: 59-60 mm SL, Ducie; observed at Henderson in 55 m. Macropharyngodon meleagris (Valenciennes, 1839). BPBM 12242, 76 mm SL, Ducie; BPBM 16611, 64 mm SL, Oeno; BPBM 16939, 2: 45-109 mm SL, Pitcairn; BPBM 17046, 4: 45-59 mm SL, Pitcairn. Novaculichthys taeniourus (Lacepéde, 1801). Observed at Oeno. Oxycheilinus unifasciatus (Streets, 1877). BPBM 16778, 233 mm SL, Pitcairn; BPBM 17129, 220 mm SL, Ducie (identfied as Cheilinus rhodochrous by Rehder and Randall, 1975). Oxycheilinus sp. An undescribed species under study by Mark Westneat, Martin F. Gomon, and the author. One was speared in 49 m at Henderson but escaped. Pseudocheilinus octotaenia Jenkins, 1901. BPBM 16541, 6: 48-69 mm SL, Oeno; BPBM 17041, 90 mm SL, Pitcairn; observed at Ducie and Henderson. Pseudocheilinus tetrataenia Schultz, 1960. BPBM 12273, 3: 23-48 mm SL, Ducie; BPBM 16503, 7: 33-47 mm SL, Oeno; BPBM 16529, 3: 26-55 mm SL, Oeno; BPBM 16884, 4: 33-47 mm SL, Pitcairn; observed at Henderson. Pseudocheilinus citrinus Randall, 1999. BPBM 16863, 59.8 mm SL (holotype), Pitcairn; BPBM 16879, 38.8 mm (paratype), Pitcairn; BPBM 17102, 65.3 mm SL (paratype), Henderson; another paratype from Pitcairn to National Science Museum, Tokyo. Pseudocheilinus ocellatus Randall, 1999. BPBM_ 16900, 39.5 mm SL (paratype), Pitcairn. Pseudojuloides atavai Randall and Randall, 1981. BPBM 16902, 115.8 mm SL (holotype), Pitcairn; paratypes from Pitcairn and Oeno to other museums; observed at Ducie. Pseudolabrus fuentesi (Regan, 1913). BPBM 15073, 10: 35-135 mm SL, Pitcairn; BPBM 16976, 10: 32-47 mm SL, Pitcairn; BPBM 17019, 10: 37-61 mm SL, Pitcairn; BPBM 17044, 68 mm SL, Pitcairn. Stethojulis bandanensis (Bleeker, 1851). BPBM 12264, 2: 80-103 mm SL, Ducie; BPBM 16500, 60 mm SL, Oeno; BPBM 16699, 65 mm SL, Pitcairn; BPBM 16802, 4: 77-82 mm SL, Pitcairn. Thalassoma heiseri Randall and Edwards, 1984. BPBM 12266, 8: 37.6-104.6 mm SL (paratypes), Ducie; BPBM 16511, 2: 57.2-89.7 mm SL (paratypes), Oeno; BPBM 16517, 7: 31.1-102 mm SL (paratypes), Oeno; BPBM 16682, 70.4 mm SL (holotype), Pitcairn; BPBM 17004, 9: 41.2-88 mm SL (paratypes), Pitcairn; BPBM 17059, 8: 38.2-116.2 mm SL (paratypes), Henderson. 23 Thalassoma lutescens (Lay and Bennett, 1839). BPBM 12265, 5: 48-86 mm SL, Ducie; BPBM 16510, 3: 81-116 mm SL, Oeno; BPBM 16674, 203 mm SL, Pitcairn; BPBM 16695, 2: 41-121 mm SL, Pitcairn; BPBM 16959, 39 mm SL, Pitcairn; BPBM 16967, 2: 42-78 mm SL, Pitcairn; BPBM 17024, 60 mm SL, Pitcairn; BPBM 17957, 5: 65-150 mm SL, Henderson; BPBM 38417, 35 mm SL, Oeno. Thalassoma purpureum (Forsskal, 1775). BPBM 13235, 240 mm SL, Pitcairn; BPBM 16811, 196 mm SL, Pitcairn; BPBM 17056, 6: 32-112 mm SL, Henderson; BPBM 17106, 2: 259-272 mm SL, Henderson; BPBM 17138, 3: 115-277 mm SL, Ducie; BPBM 28929, 2: 83-100 mm SL, Pitcairn; observed at Oeno. Thalassoma trilobatum (Lacepéde, 1801). BPBM 17020, 2: 125-153 mm SL, Pitcairn; observed at Oeno and Henderson. Wetmorella nigropinnata (Seale, 1901). BPBM 16544, 2: 53-58 mm SL, Oeno. Xyrichtys pavo Valenciennes, 1839. BPBM 16832, 266 mm SL, Pitcairn; observed at Henderson (Irving et al, 1995). SCARIDAE (PARROTFISHES) Calotomus carolinus (Valenciennes, 1839). BPBM 16658, 398 mm SL (head and caudal fin saved), Pitcairn. Chlorurus frontalis (Valenciennes, 1840). BPBM 16596, 342 mm SL, Oeno; BPBM 16632, 508 mm SL (saved head), Pitcairn; BPBM 16808, 6: 22-142 mm SL, Pitcairn; BPBM 16837, 595 mm SL (saved head), Pitcairn. Chlorurus microrhinos (Bleeker, 1854). BPBM 16598, 460 mm SL (saved head), Oeno; BPBM 16825, 404 mm SL (saved head and caudal fin), Pitcairn; BPBM 17115, 312 mm SL, Ducie; BPBM 17116, 325 mm SL, Ducie; observed at Henderson. Chlorurus sordidus (Forsskal, 1775). BPBM 17134, 214 mm SL, Ducie; BPBM 38416, 79 mm SL, Oeno. Leptoscarus vaigiensis (Quoy and Gaimard, 1824). BPBM 16815, 258 mm SL, Pitcairn. Scarus altipinnis (Steindachner, 1879). BPBM 16836, 2: 426-527 mm SL (saved heads and caudal fins), Pitcairn. Scarus forsteni (Bleeker, 1861). BPBM 16597, 361 mm SL, Oeno; BPBM 16704, 335 mm SL, Pitcairn; BPBM 16839, 392 mm SL, Pitcairn; BPBM 17123, 325 mm SL, Ducie; BPBM 17152, 298 mm SL, Ducie; observed at Henderson. Scarus frenatus Lacepede, 1802. BPBM 17156, 370 mm SL, Ducie; BPBM 17160, 222 mm SL, Ducie. Scarus ghobban Forsskal, 1775. Observed at Ducie. 24 Scarus longipinnis Randall and Choat, 1980. BPBM 16648, 280 mm SL (paratype), Pitcairn; BPBM 16656, 209 mm SL (paratype), Pitcairn; BPBM 16680, 274 mm SL (holotype), Pitcairn; BPBM 16763, 170 mm SL (paratype), Pitcairn; BPBM 16773, 211 mm SL (paratype), Pitcairn; observed at Henderson. PINGUIPEDIDAE (SAND PERCHES) Parapercis millepunctata (Giinther, 1860). BPBM 16454, 134 mm SL, Pitcairn; BPBM 17029, 73 mm SL, Pitcairn. Parapercis multiplicata Randall, 1984. BPBM 16451, 63 mm SL, Pitcairn. (Fig. 30). Parapercis schauinslandiu (Steindachner, 1900). BPBM 16453, 71 mm SL, Pitcairn. TRIPTERYGIIDAE (TRIPLEFINS) Enneapterygius ornatus Fricke, 1997. BPBM 17081, 21.8 mm SL (holotype), Henderson; BPBM uncat., 16: 19.5-24.4 mm SL (paratypes), Henderson. Enneapterygius pyramis Fricke, 1994. BPBM 16486, 20.7 mm SL, Oeno; BPBM 16537, 2: 13.8-15 mm SL, Oeno; BPBM 16916, 6: 21.2-24.6 mm SL, Pitcairn; BPBM 16923, 12: 11.4-26.2 mm SL, Pitcairn; BPBM 16948, 4: 17.4-21.7 mm SL, Pitcairn; BPBM 16957, 5: 13.7-25.9 mm SL, Pitcairn; BPBM 16972, 14: 11.8-21.8 mm SL, Pitcairn; BPBM 16985, 7: 15.2-23.9 mm SL, Pitcairn; BPBM 16998, 23 mm SL, Pitcairn; BPBM 17002, 6: 21.2- 25.6 mm SL, Pitcairn; BPBM 17003, 3: 13.6-23 mm SL, Pitcairn; BPBM 17026, 3: 21.1-21.6 mm SL, Pitcairn; BPBM 17028, 2: 23.4-24.1 mm SL, Pitcairn. (Fig. 31). Norfolkia thomasi Whitley, 1964. BPBM 16897, 34 mm SL, Pitcairn. CREEDIDAE (SAND BURROWERS) Chalixodytes tauensis Schultz, 1943. BPBM 17025, 2: 35-38 mm SL, Pitcairn; BPBM 29655, 3: 35-38 mm SL, Pitcairn. Crystallodytes cookei enderburyensis Schultz, 1943. BPBM 12269, 2: 35-42 mm SL, Ducie; BPBM 16871, 41 mm SL, Pitcairn; BPBM 16979, 40 mm SL, Pitcairn; BPBM 16992, 11: 19-41 mm SL, Pitcairn; BPBM 17098, 11: 35-52 mm SL, Pitcairn. Limnichthys donaldsoni Schultz, 1960. BPBM 16886, 2: 18-19 mm SL; BPBM 16958, 2: 24-27 mm SL, Pitcairn; BPBM 17086, 19 mm SL, Henderson. AMMODYTIDAE (SAND LANCES) Ammodytes sp. An undescribed species from Pitcairn under study by Bruce B. Collette and the author. BLENNIIDAE (BLENNIES) Alticus sp. BPBM 16438, 64 mm SL, Pitcairn; BPBM 16801, 6: 28-66 mm SL, Pitcairn; BPBM 16866, 3: 43-72 mm SL, Pitcairn; BPBM 17016, 28 mm SL, 25 Pitcairn; BPBM 17097, 34: 22-61 mm SL, Henderson; BPBM 36430, 7: 36- 51 mm SL, Henderson; MCZ 35569, 3: 39-55 mm, collected at Pitcairn by H. Schroeder, 14 December 1939; examined by the author at the Museum of Comparative Zoology, Harvard University. A new species that will be described in a revision of the genus by Jeffrey T. Williams. Blenniella gibbifrons (Quoy and Gaimard, 1824). BPBM 16733, 4: 43-73 mm SL, Pitcairn; BPBM 17010, 10: 39-84 mm SL; BPBM 17095, 16: 32-69 mm SL, Henderson; BPBM 17147, 11: 39-64 mm SL, Ducie. Blenniella paula (Bryan and Herre, 1903). BPBM 17128, 2: 52-60 mm SL, Ducie. Cirripectes alboapicalis (Ogilby, 1899). BPBM 12267, 2: 21-38 mm SL, Ducie; BPBM 16482, 2: 38-42 mm SL, Oeno; BPBM 16730, 2: 70-82 mm SL, Pitcairn; BPBM 16974, 14: 35-58 mm SL, Pitcairn; BPBM 16996, 37 mm SL, Pitcairn; BPBM 17096, 15: 41-73 mm SL, Henderson; BPBM 30983, 30 mm, Henderson. (Fig. 32). Cirripectes quagga (Fowler and Ball, 1924). BPBM 17070, 10: 29-46 mm SL, Henderson. (Fig. 33). Cirripectes variolosus (Valenciennes, 1836). BPBM 12270, 4: 28-61 mm SL, Ducie; BPBM 16928, 2: 42-43 mm SL, Pitcairn; BPBM 16941, 2: 25-53 mm SL, Pitcairn; BPBM 17009, 40 mm SL, Pitcairn; BPBM 17079, 4: 25-29 mm, Henderson; BPBM 31010, 5: 39-47 mm SL, Pitcairn. Entomacrodus caudofasciatus (Regan, 1909). BPBM 13749. 2: 32-36 mm SL, Ducie; BPBM 13753, 9: 40.5-55 mm SL, Pitcairn; BPBM 17094, 6: 29-55 mm SL, Henderson. Entomacrodus niuafoouensis (Fowler, 1932). BPBM 13745, 2: 65-80 mm SL, Henderson; BPBM 13756, 7: 32-97 mm SL, Pitcairn; BPBM 13757, 13: 18- 21 mm SL, Pitcairn. (Fig. 34). Entomacrodus rofeni Springer, 1967. BPBM 13746, 48 mm SL, Henderson; BPBM 13750, 5: 49-52 mm SL, Ducie; BPBM 13751, 4: 46-49 mm SL, Ducie; BPBM 17092, 20: 34-53 mm SL, Henderson. (Fig. 35). Entomacrodus sealet Bryan and Herre, 1903. BPBM 13754, 10: 41.6-61.2 mm SL, Pitcairn; BPBM 17091, 10; 19-96 mm SL, Henderson. Entomacrodus striatus (Valenciennes, 1836). BPBM 13748, 44 mm SL, Ducie; BPBM 16437, 63 mm SL, Pitcairn; BPBM 17018, 69: 28-104 mm SL, Pitcairn; BPBM 36431, 52 mm SL, Henderson. Exallias brevis (Kner, 1968). BPBM 17062, 92 mm SL, Henderson. Istiblennius edentulus (Bloch and Schneider, 1801). BPBM 17021, 3: 109-136 mm SL, Pitcairn. 26 Plagiotremus tapeinosoma (Bleeker, 1857). BPBM 16672, 45 mm SL, Pitcairn; BPBM 16514, 88 mm SL, Oeno; BPBM 17104, 44 mm SL, Henderson. Praealticus caesius (Seale, 1906). BPBM 16729, 23: 18-67 mm SL, Pitcairn; BPBM 16800, 6: 43-59 mm SL, Pitcairn. (Fig. 36). Rhabdoblennius ellipes (Jordan and Starks, 1906). BPBM 13752, 11: 26-39 mm SL, Ducie; BPBM 16734, 3: 31-33 mm SL, Pitcairn; BPBM 16903, 3: 30-39 mm SL, Pitcairn; BPBM 17080, 17: 26-41 mm SL, Henderson. Identified as R. rhabdotrachelus (Fowler and Ball) by Irving et al. (1995), but this appears to be a junior synonym of R. ellipes. (Fig. 37). Stanulus seychellensis Smith, 1959. BPBM 13747, 29 mm SL, Henderson. GOBIESOCIDAE (CLINGFISHES) Lepadichthys frenatus Waite, 1904. BPBM 16921, 5: 13-39 mm SL, Pitcairn; BPBM 38419, 8: 19-35 mm SL, Oeno. Pherallodus indicus (Weber, 1913). BPBM 17071, 10.5 mm SL, Henderson. Identified by John C. Briggs. CALLIONYMIDAE (DRAGONETS) Synchiropus ocellatus (Pallas, 1770). BPBM 16473, 42 mm SL, Oeno; BPBM 16946, 24 mm SL, Pitcairn. GOBIIDAE (GOBIES) Bathygobius cocosensis (Bleeker, 1854). BPBM 16874, 41 mm SL, Pitcairn. Bathygobius cotticeps (Steindachner, 1880). BPBM 17008, 35 mm SL, Pitcairn. Eviota albolineata Jewett and Lachner, 1983. BPBM 17142, 17.5 mm SL, Ducie. Eviota saipanensis Fowler, 1945. BPBM 17076, 18 mm SL, Henderson. Gnatholepis cauerensis (Bleeker, 1853). BPBM 16532, 7: 25-49 mm SL, Oeno; BPBM 16538, 10: 22-50 mm SL, Oeno; BPBM 16846, 2: 41-44 mm SL, Pitcairn; BPBM 16880, 33 mm SL, Pitcairn; BPBM 38423, 2: 20-22 mm SL, Oeno. These specimens, along with ones from Rapa, Austral Islands, and Rarotonga are being described as a new subspecies by Randall and Greenfield (MS). Gobiodon sp. BPBM 17090, 3: 19-27 mm SL, Henderson. These specimens are characterised by uniform brown colour in alcohol (finely speckled with dark brown under magnification), dorsal rays VI-I,9; last membrane of spinous dorsal fin linked to first spine of second dorsal fin about one-fifth spine length above its base; anal rays [,8; pectoral rays 18; gill rakers 2 + 7; a broad groove between interopercle and isthmus; gill opening equals full height of pectoral-fin base. Gobiodon sp. of BPBM 17227, 17 mm SL, reported by Randall et al. 27 (1990) appears to be the same species. A revision of this genus is needed. Hetereleotris sp. BPBM 38424 and 38425 from Oeno on loan to Douglass F. Hoese since 1983. Paragobiodon sp. BPBM 16498, 3: 15-21 mm SL, Oeno; BPBM 16526, 3: 21-28 mm SL, Oeno; BPBM 16913, 2: 26-27 mm SL, Pitcairn. These specimens have numerous large fleshy papillae on the cheek and ventrally on the head, small slender papillae dorsally on the head and nape, rows of papillae along the membranes of the ventral surface of the pelvic disc, no scales ventrally on abdomen, and 23 pectoral rays; they are brown in preservative, the head paler than body, and the median fins darker. BPBM 13545, 2: 22-23 mm SL, from Mangareva may be the same species. Paragobiodon is also in need of revision. Priolepis squamogena Winterbottom and Burridge, 1989. BPBM 12249, 5: 29- 38 mm SL, Ducie; BPBM 16911, 27 mm SL, Pitcairn; BPBM 16981, 30 mm SL, Pitcairn; BPBM 16989, 3: 32-41 mm SL, Pitcairn. Reported from Ducie by Rehder and Randall (1975: 24) as Quisquilius cinctus (Regan). Priolepis semidoliatus (Valenciennes, 1830). BPBM 12260, 6: 12-22 mm SL, Ducie; BPBM 16994, 6: 17-19 mm SL, Pitcairn; BPBM 17141, 13 mm, Ducie. Trimmatom sp. BPBM 16915, 11 mm SL, Pitcairn; BPBM 16935, 2: 11-16 mm SL, Pitcairn; BPBM 17012, 17 mm SL, Pitcairn; BPBM 17072, 13 mm SL, Henderson; BPBM 17143, 11 mm SL, Ducie; BPBM 38433, 2: 9-14 mm SL, Oeno. An undescribed species of the eviotops complex under study by Richard Winterbottom. MICRODESMIDAE (DARTFISHES AND WORMFISHES) Nemateleotris magnifica Fowler, 1938. BPBM 16612, 5: 61-64 mm SL, Oeno; BPBM 16908, 48 mm SL, Pitcairn; observed at Henderson (Irving et al., 1995) Ptereleotris evides (Jordan and Hubbs, 1925). Observed at Oeno. ACANTHURIDAE (SURGEONFISHES) Acanthurus achilles (Shaw, 1803). BPBM 17158, 170 mm SL, Ducie; underwater photo taken at Henderson. Acanthurus guttatus Bloch and Schneider, 1801. BPBM 13239, 150 mm SL, Pitcairn; BPBM 17101, 190 mm SL, Henderson. Acanthurus leucopareius (Jenkins, 1903). BPBM 12251, 4: 63-78 mm SL, Ducie; BPBM 16583, 2: 141-172 mm SL, Oeno; BPBM 16688, 61 mm SL, Pitcairn; BPBM 16720, 3: 85-124 mm SL, Pitcairn; BPBM 17066, 3: 72-109 28 mm SL, Henderson. Acanthurus nigrofuscus (Forsskal, 1775). BPBM 13244, 130 mm SL, Pitcairn; BPBM 17064, 84 mm SL, Henderson. Acanthurus nigroris Valenciennes, 1835. BPBM 12279, 149 mm SL, Henderson. Acanthurus nubilus (Fowler and Bean, 1929). BPBM 16710, 189 mm SL, Pitcairn; BPBM 17155, 200 mm SL, Ducie; observed at Henderson. (Fig. 38). Acanthurus thompsoni (Fowler, 1923). BPBM 16582, 2: 164-167 mm SL, Oeno; BPBM 16771, 114 mm SL, Pitcairn; observed at Ducie and Henderson. Acanthurus triostegus (Linnaeus, 1758). BPBM 12239, 149 mm SL, Ducie; BPBM 16569, 2: 64-141 mm SL, Oeno; BPBM 16668, 152 mm SL, Pitcairn; BPBM 16731, 6: 70-86 mm SL, Pitcairn; observed at Henderson (Irving et al., 1995). Ctenochaetus hawatiensis Randall, 1955. BPBM 16681, 214 mm SL, Pitcairn; BPBM 17164, 220 mm SL, Ducie; observed at Henderson. Ctenochaetus striatus (Quoy and Gaimard, 1825). BPBM 16584, 9: 59-159 mm SL, Oeno; BPBM 16751, 210 mm SL, Pitcairn; observed at Henderson (Irving et al., 1995). Ctenochaetus strigosus (Bennett, 1828). BPBM 13249, 112 mm SL, Pitcairn; BPBM 16570, 4: 99-126 mm SL, Oeno; BPBM 17135, 72 mm SL, Ducie; observed at Henderson. Naso brevirostris (Valenciennes, 1835). Observed at Ducie and Henderson. Naso caesius Randall and Bell, 1992. BPBM 16813, 407 mm SL (saved head and caudal fin), Pitcairn; observed at Henderson. Naso hexacanthus (Bleeker, 1855). Observed at Pitcairn, Ducie, and Henderson. Naso lituratus (Forster in Bloch and Schneider, 1801). BPBM 16616, 292 mm SL (saved head), Oeno; BPBM 16780, 269 mm SL, Pitcairn. Naso unicornis (Forsskal, 1775). BPBM 16595, 303 mm SL, Oeno; BPBM 16776, 536 mm SL (saved head and caudal fin with spines), Pitcairn; observed at Henderson (Irving et al., 1995). Zebrasoma rostratum (Giinther, 1873). BPBM 16711, 164 mm SL, Pitcairn; BPBM 17136, 182 mm SL, Ducie. Zebrasoma scopas (Cuvier, 1829). BPBM 16709, 175 mm SL, Pitcairn; observed at Henderson. Zebrasoma veliferum (Bloch, 1797). BPBM 16827, 190 mm SL, Pitcairn; observed at Henderson (Irving et al., 1995). ZANCLIDAE (MOORISH IDOL FAMILY) 29 Zanclus cornutus (Linnaeus, 1758). BPBM 17136, 147 mm SL, Oeno; observed at the other three islands. SIGANIDAE (RABBITFISHES) Siganus argenteus (Quoy and Gaimard, 1825). BPBM 17931, 310 mm SL, Pitcairn. TRICHIURIDAE (CUTLASSFISHES) Benthodesmis sp. BPBM 36423, 113 mm SL, Henderson; regurgitated by a juvenile masked booby. SPHYRAENIDAE (BARRACUDAS) Sphyraena helleri Jenkins, 1901. BPBM 16600, 603 mm SL, Oeno; BPBM 16760, 540 mm (saved head and caudal fin), Pitcairn; BPBM 36427, 540 mm SL, Henderson (reported as S. novaehollandae by Irving et al., 1995). GEMPYLIDAE (SNAKE MACKERELS) Gempylus sp. AMNH 8261, 193 mm SL, Ducie, from a white tern. Reported by Nichols (1923) as Lemnisoma thyrsitoides Lesson. The specimen was borrowed from the American Museum of Natural History by the author and reidentified only to genus (noted as having 31 dorsal spines). SCOMBRIDAE (TUNAS AND MACKERELS) Acanthocybium solandri (Cuvier, 1831). BPBM 16829, 1290 mm SL (saved head), Pitcairn. Katsuwonus pelamis (Linnaeus, 1758). On 2-5 February 1980 a tuna survey of the South Pacific Commission in Pitcairn waters resulted in the capture of 11 skipjack tuna. Gymnosarda unicolor (Riippell, 1836). BPBM 16966, 1080 mm FL (saved head), Oeno; observed at Pitcairn and Henderson. Thunnus albacares (Bonnaterre, 1788). BPBM 16779, 1445 mm FL (saved head and fins), Pitcairn. Thunnus obesus (Lowe, 1839). The above-mentioned tuna survey of 1980 resulted in 321 yellowfin tuna and two bigeye tunas. BOTHIDAE (LEFT EYE FLOUNDERS) Bothus mancus (Broussonet, 1782). BPBM 16571, 205 mm SL, Oeno; BPBM 16662, 142 mm SL, Pitcairn; BPBM 17013, 2: 74-75 mm SL, Ducie; BPBM 17149, 184 mm SL, Ducie; BPBM 36428, 225 mm SL, Henderson. SAMARIDAE (SLENDER FLOUNDERS) Samariscus triocellatus Woods in Schultz, 1953. BPBM 16608, 40 mm SL, Oeno. BALISTIDAE (TRIGGERFISHES) 30 Balistoides viridescens (Bloch and Schneider, 1801). Observed at Pitcairn and Henderson. Pseudobalistes fuscus (Bloch and Schneider, 1801). BPBM 16599, 374 mm SL, Oeno; observed at Pitcairn and Ducie. Rhinecanthus aculeatus (Linnaeus, 1758). BPBM 16565, 160 mm SL, Oeno; observed at Henderson (Irving et al., 1995). Rhinecanthus lunula Randall and Steene, 1983. BPBM 17177, 218 mm SL (paratype), Ducie; observed at Pitcairn and Henderson. Rhinecanthus rectangulus (Bloch and Schneider, 1801). BPBM 12241, 123 mm SL, Ducie; BPBM 12280, 174 mm SL, Henderson; BPBM 16667, 152 mm SL, Pitcairn; observed at Oeno. Sufflamen bursa (Bloch and Schneider, 1801). BPBM 13255, 180 mm SL, Pitcairn; BPBM 16557, 122 mm SL, Oeno; BPBM 16769, 162 mm SL, Pitcairn; BPBM 17157, 158 mm SL, Ducie; observed at Henderson. Sufflamen fraenatus (Latreille, 1804). BPBM 16575, 221 mm SL, Oeno; observed at Henderson. Xanthichthys mento (Jordan and Gilbert, 1882). BPBM 13256, 2: 130-158 mm SL, Pitcairn; BPBM 16822, 5: 130-155 mm SL, Pitcairn; BPBM 16834, 4: 142-162 mm SL, Pitcairn. MONACANTHIDAE (FILEFISHES) Aluterus scriptus (Osbeck, 1765). Observed at Pitcairn and Henderson. Cantherhines dumerilit (Hollard, 1854). BPBM 16669, 210 mm SL, Pitcairn; BPBM 17167, 273 mm SL, Ducie; observed at Henderson. Cantherines pardalis (Riippell, 1837). BPBM 12246, 130 mm SL, Ducie; BPBM 16675, 155 mm SL, Pitcairn; observed at Henderson. OSTRACIIDAE (TRUNKFISHES) Ostracion meleagris Shaw, 1796. Observed at Henderson (Irving et al., 1995). TETRAODONTIDAE (PUFFERS) Arothron meleagris (Lacepede, 1798). BPBM 16564, 170 mm SL, Oeno; BPBM 16748, 195 mm SL, Pitcairn; BPBM 17121, 233 mm SL, Ducie; observed at Henderson. Canthigaster janthinoptera (Bleeker, 1855). BPBM 11171, 3: 42.5-54 mm SL, Oeno; BPBM 11182, 37 mm SL, Oeno; BPBM 11187, 6: 38.5-51 mm SL, Oeno; BPBM 11209, 20 mm, Pitcairn. DIODONTIDAE (PORCUPINEFISHES) Diodon holocanthus Linnaeus, 1758. BPBM 13251, 148 mm SL, Pitcairn; BPBM 16445, 123 mm SL, Pitcairn; BPBM 16459, 2: 144-168 mm SL, Oeno; observed at Henderson (Irving et al., 1995). Diodon hystrix Linnaeus, 1758. BPBM 16631, 275 mm SL, Pitcairn; BPBM 31 16717, 2: 294-380 mm SL, Pitcairn; BPBM 16821, 278 mm SL, Pitcairn; one of 380 mm SL collected at Ducie but not saved; observed at Oeno. MOLIDAE (MOLAS) Ranzania laevis (Pennant, 1776). BPBM 36425, 96 mm SL and BPBM 36426, 113 mm SL. Both dropped on land at Henderson by fairy terns. DISCUSSION The most common fishes observed at diving depths of about 10-20 m at Pitcairn were the wrasses Thalassoma lutescens and Coris sp., the surgeonfish Acanthurus leucopareius, the damselfishes Chrysiptera galba and Stegastes fasicolatus, and the sea chub Kyphosus bigibbus. The last-mentioned was not uniformly distributed but occurred in large aggregations, usually near a rocky area with ledges and caves. This fish is the most important food fish for the islanders who catch it in large numbers with hook and line. An occasional individual is entirely bright yellow; rarely one may see an all-white one. There are also fish of mixed yellow and brown, white and brown, and most rarely, white and yellow. Aggregations of Chromis bami and C. pamae were locally abundant. In the shallower water the wrasse Pseudolabrus fuentesit was most common where the bottom was thickly covered with brown algae, and Thalassoma heiseri on hard substratum where there was little algal cover. In their respective families, the most common species as viewed by divers were Chaetodon smithi of the butterflyfishes, Epinephelus fasciatus of the groupers, Scarus longipinnis of the parrotfishes, and Myripristis berndti of the squirrelfishes and soldierfishes. Jacks were not common at Pitcairn compared to the other three uninhabited islands, probably because of fishing pressure. Those species most often seen were Caranx lugubris, Pseudocaranx dentex, and Carangoides orthogrammus. The most common fishes observed at Ducie Atoll were Kyphosus bigibbus, Thalassoma lutescens, T. heiseri, Chlorurus microrhinos, Acanthurus leucopareius, Chrysiptera galba, Stegastes emeryi, and Caranx lugubris. One can count 16 solitary individuals of C. lugubris in the underwater photograph of Fig. 28 in Rehder and Randall (1975). This brazen fish often approached the divers closely; one struck the author with its tail as it swam past, and some were fed by hand with pieces of octopus. Chaetodon ulietensis was the most common of the butterflyfishes (Figs. 24 and 25 in Rehder and Randall, 1975), Epinephelus fasciatus and E. tauvina the most common of the groupers, and Parupeneus bifasciatus and P. cyclostomus the most common of the goatfishes. No 32 individuals of the two unicornfishes Naso lituratus and N. unicornis were seen at Ducie; this might be related to the paucity of brown algae at the atoll. The fish fauna of the Pitcairn Islands is impoverished compared to that of the archipelagos to the west. The total number of fishes listed in the present report is 348, of which 13 are pelagic species. Randall (1985) listed 593 species of shore fishes for the Society Islands alone. The reason for the relatively few Pitcairn species of fishes is undoubtedly related to the isolation of the islands, their more southern location, their small size, and the lack of some habitats such as estuaries, mangroves, and seagrass beds. Still, the Pitcairn fish fauna is far larger than that of the very remote Easter Island, with 126 shore species. As would be expected, the fish fauna of the Pitcairn Islands is very similar to that of the southeastern islands of the Tuamotu Archipelago; however, there are some species that were seen at the Gambier Group of the Tuamotus (our field work at Mangareva and Temoe Atoll) but not sighted in the Pitcairn Islands. Tuamotu fishes that are apparently absent from the Pitcairn Islands include: Gymnothorax pictus, Epinephelus polyphekadion, Pseudanthias pascalus, Lutjanus fulvus, Parupeneus_ barberinus, Hoplolatilus starcki, Ellochelon vaigiensis , Plectroglyphidodon lacrymatus, Pomacentrus coelestis, Bodianus perditio, Epibulus insidiator, Cetoscarus bicolor, Hipposcarus longiceps, Sphyraena genie, Ostracion cubicus, Canthigaster solandri, and C. valentint. Of the 335 shore fishes of the Pitcairn Islands, 284 are tropical Indo-Pacific species or ones that are wide-ranging in the central and western Pacific. Five species of fishes are presently known only from the Pitcairn Islands: Sargocentron megalops, Hemitaurichthys multispinosus, Ammodytes sp., Enneapterygius ornatus, and Alticus sp. The following 21 species of Pitcairn fishes occur only in the South Pacific below a latitude of 20°S: Gymnothorax australicola, G. nasuta, Myripristis tiki, Sargocentron hormion, Cosmocampus howensis, Pseudogramma australis, Amblycirrhitus wilhelmi, Chaetodon flavirostris, Chaetodon smith, Centropyge hotumatua, Genicanthus spinus, Chromis bamit, C. pamae, Chrysiptera galba, Anampses femininus, Pseudocheilinus citrinus, Pseudolabrus fuentesi, Scarus longipinnis, Cirripectes alboapicalis, Praealticus caesius, and Gobiodon sp. Some of these are wide-ranging across much of the South Pacific. Gymnothorax australicola, Cosmocampus howensis, and Anampses femininus occur from Easter Island to Lord Howe Island or New South Wales. Sixteen other Pitcairn fishes are confined to the Southern Hemisphere at latitudes greater than 14° S: Myripristis randalli, Scorpaenopsis sp., Epinephelus tuamotuensis, Pseudanthias mooreanus, Paracirrhites nisus, Cercamia cladara, Pomachromis fuscidorsalis, Stegastes emeryi, Cirrhilabrus scottorum, Coris sp., 33 Oxycheilinus sp., Thalassoma heisert, Entomacrodus rofeni, Paragobiodon sp., Trimmatom sp., and Rhinecanthus lunula. All but the last-mentioned are restricted to the Pitcairn Islands and the islands of French Polynesia except the Marquesas. Twelve species of Pitcairn fishes are antitropical or antiequatorial in distribution based on present knowledge: Gymnothorax eurostus, Saurida flamma, Synodus capricornis, Iracundus signifer, Sebastapistes tinkhami, Pseudocaranx dentex, Sertola lalandi, Pseudojuloides atavai, Lepadichthys frenatus, Norfolkia thomasi, Acanthurus leucopareius, and Xanthichthys mento. ACKNOWLEDGEMENTS The expedition on “Westward” was supported principally by a grant from the National Geographic Society to the Bishop Museum. The vessel was made available by the Oceanic Foundation in Hawaii, as well as half her operating costs. The Hawaii Institute of Marine Biology of the University of Hawaii loaned a small boat and diving gear. Assistance in the collection of fishes was provided by J. David Bryant, Dean B. Cannoy, R. Richard Costello, Dr. Guy Haywood, his son the late James R. Haywood, Denis N. Hewett, the late Rhett M. McNair, Akihiko Sinoto, and several of the Pitcairn Islanders, in particular Stephen Christian and Noggie Young. Robert A. Irving donated Henderson Island fish specimens to the Bishop Museum. Special thanks are due the friendly people of Pitcairn for their gracious hospitality and help in innumerable ways. I am grateful to Eugenia B. Bohlke, John C. Briggs, William N. Eschmeyer, Thomas H. Fraser, David W. Greenfield, Walter Ivantsoff, John E. McCosker, Randall D. Mooi, David G. Smith, Victor G. Springer, Kenneth A. Tighe, Jeffrey T. Williams, and Richard Winterbottom for their aid in the identification of Pitcairn fishes. Also thanks to Roy T. Tsuda who identified marine algae from Pitcairn, and Lars-Ake Gothesson for some records of Pitcairn fishes after the 1970-71 visit on “Westward”. The manuscript was reviewed by Robert F. Myers. LITERATURE CITED ALLEN, G.R. & J.E. RANDALL. 1974. -- Five new species and a new genus of damselfishes (Family Pomacentridae) from the South Pacific Ocean. Tropical Fish Hobbyist 21 (9): 36-46, 48-49. 34 CASTLE, P.H.J. 1968. -- A contribution to a revision of the moringuid eels. Special Publication of the Department of Ichthyology, Rhodes University, Grahamstown 3: 1-29. CRESSEY R. & J.E. RANDALL. 1978. -- Synodus capricornis, a new lizardfish from Easter and Pitcairn Islands. Proceedings of the Biological Society of Washington 91 (3): 767-774, DISALVO, L.H., J-E. RANDALL, & A. CEA. -- 1988. Ecological reconnaissance of the Easter Island sublittoral environment. National Geographic Research 4 (4): 451-473. FRICKE, R. 1997. --Tripterygiid Fishes of the Western and Central Pacific (Teleostei). Koeltz Scientific Books, Koenigstein, Germany, ix + 607 pp. GREENFIELD, D.W. 1974. -- A revision of the squirrelfish genus Myripristis Cuvier (Pisces: Holocentridae). Natural History Museum Los Angeles County Science Bulletin 19: 1-54. GOSLINE, W.A. & D.W. STRASBURG. 1956. -- The Hawaiian fishes of the family Moringuidae: another eel problem. Copeia 1956 (1): 9-18. HERALD, E.W. & J.E. RANDALL. 1972. -- Five new Indo-Pacific pipefishes. Proceedings of the California Academy of Sciences 39 (11): 121-140. IRVING, R.A.,, J. JAMIESON & J.E. RANDALL. 1995. -- Initial checklist of fishes from Henderson Island, Pitcairn Group. Biological Journal of the Linnaean Society 56: 329-338. LAVENBERG, RJ. 1992. -- A new moray eel (Muraenidae: Gymnothorax) from oceanic islands of the South Pacific. Pacific Science 46 (1): 58-67. MCCOSKER, J.E. & J.E. RANDALL. 1977. - Three new species of Indo-Pacific moray eels (Pisces: Muraenidae). Proceedings of the California Academy of Sciences 41 (3): 161-168. NICHOLS, J.T. 1923. -- Two new fishes from the Pacific Ocean. American Museum Novitates 94: 1-3. PAULAY, G. 1989. -- Marine invertebrates of the Pitcairn Islands: species composition and biogeography of corals, molluscs, and echinoderms. Atoll Res. Bull. 326: 1-28. RANDALL, H.A. & G.R. ALLEN. 1977. -- A revision of the damselfish genus Dascyllus (Pomacentridae) with description of a new species. Records of the Australian Museum 31 (9): 349-385. RANDALL, J.E. 1972. — A revision of the labrid fish genus Anampses. Micronesica 8 (1 & 2): 151-190. RANDALL, J.E. 1973. - Expedition to Pitcairn. Oceans 6: 12-21. RANDALL, J.E. 1975a. - Three new butterflyfishes (Chaetodontidae) from southeast Oceania. UO 25: 12-22. 35 RANDALL, J.E. 1975b. — A revision of the Indo-Pacific angelfish genus Genicanthus, with descriptions of three new species. Bulletin of Marine Science 25 (3): 393-421. RANDALL, J.-E. 1978. -- Marine biological and archaeological expedition to southeast Oceania. National Geographic Research Reports 1969: 473-495. RANDALL, J.E. 1979. - A review of the serranid fish genus Anthias of the Hawaiian Islands, with description of two new species. Contributions in Science Natural History Museum of Los Angeles County 302: 1-13. RANDALL, J.E. 1980. — Revision of the fish genus Plectranthias (Serranidae: Anthiinae) with descriptions of 13 new species. Micronesica 16 (1): 101-187. RANDALL, J.E. 1985. -- Fishes, pp. 462-481 in B. Delesalle, R. Galzin and B. Salvat (eds.). Fifth International Coral Reef Congress, Tahiti, 27 May - 1 June: “French Polynesian Coral Reefs.” RANDALL, J.E. 1998. -- Revision of the Indo-Pacific squirrelfishes (Beryciformes: Holocentridae: Holocentrinae) of the genus Sargocentron, with descriptions of four new species. Indo-Pacific Fishes 27: 1-105. RANDALL, J.E. 1999. -- Revision of the Indo-Pacific labrid fishes of the genus Pseudocheilinus, with descriptions of three new species. Indo-Pacific Fishes 28: 1-34. RANDALL, J.E. & C.C. BALDWIN. 1997. -- Revision of the serranid fishes of the subtribe Pseudogrammina, with descriptions of five new species. Indo- Pacific Fishes 26: 1-56. RANDALL, J.E. & D.K. CALDWELL. 1973. -- A new butterflyfish of the genus Chaetodon and a new angelfish of the genus Centropyge from Easter Island. Contributions in Science Natural History Museum of Los Angeles County 237: 1-11. RANDALL, J.E. & A. CEA EGANA. 1984. -- Native names of Easter Island fishes, with comments on the origin of the Rapanui people. Occasional Papers of the Bernice P. Bishop Museum 25: 1-16. RANDALL, J.E. & J.H. CHOAT. 1980. -- Two new parrotfishes of the genus Scarus from the central and south Pacific, with further examples of sexual dichromatism. Zoological Journal of the Linnaean Society 70 (4): 383-419. RANDALL, J.E. & A. EDWARDS. 1984. -- A new labrid fish of the genus Thalassoma from the Pitcairn Group, with a review of related Indo-Pacific species. Journal of Aquariculture and Aquatic Sciences 4 (2): 13-32. RANDALL, J.E. & J.E. MCCOSKER. 1992. - Two new damselfishes of the genus Chromis (Perciformes: Pomacentridae) from the South Pacific. Proceedings of the California Academy of Sciences 47 (12): 329-337. RANDALL, J.E. & R.M. PYLE. 1989. -- Cirrhilabrus scottorum, a new labrid fish from the South Pacific Ocean. Revue Frangaise d’Aquariologie (1988) 15 (4): 36 113-118. RANDALL, J.E. & H.A. RANDALL. 1981. -- A revision of the labrid fish genus Pseudojuloides, with descriptions of five new species. Pacific Science 35 (1): 51-74. RANDALL, J.E. & C.L. SMITH. 1988. -- Two new species and a new genus of cardinalfishes (Perciformes: Apogonidae) from Rapa, South Pacific Ocean. American Museum Novitates 2926: 1-9. RANDALL, J-E., C.L. SMITH & M.N. FEINBERG. 1990. -- Report on fish collections from Rapa, French Polynesia. American Museum Novitates 2966: 1- 44. REHDER, H.A & J.E. RANDALL. 1975. -- Ducie Atoll: its history, physiography and biota. Atoll Research Bulletin 183: 1-42. Figure 6. Chaetodon flavirostris, Ducie Atoll. Figure 7. The author emerging from a cave at Oeno Atoll with 2 Whitetip Reef Sharks (Triaenodon obesus) collected by spear and powerhead (Dean B. Cannoy). Figure 8. Gymnothorax eurostus, BPBM 16738, 492 mm TL, Pitcairn tidepool. Figure 10. Gymnothorax kidako, BPBM 16906, 286 mm TL, Pitcairn, 37 m. Figure 12. Uropterygius kamar, BPBM 16471, 150 mm TL, Oeno Atoll. Figure 14. Uropterygius xanthopterus, BPBM 16460, 622 mm TL, Oeno lagoon. Figure 15. Apterichtus sp., BPBM 16506, 324 mm TL, Oeno Atoll. Figure 16. Muraenichthys laticaudata, BPBM 16732, 133 mm TL, Pitcairn tidepool. Figure 19. Parascorpaena mcadamsi, BPBM 11195, 42 mm SL, Pitcairn. Figure 20. Scorpaenodes hirsutus, BPBM 11204, 40 mm SL, Pitcairn. Figure 22. Caracanthus maculatus, BPBM 16630, 40 mm SL, Oeno Atoll. Figure 24. Epinephelus lanceolatus, BPBM 33914, 1450 mm SL, 148 kg, Henderson Island. Figure 25. Seriola lalandi, BPBM 16965, 915 mm SL, 8.6 kg, Pitcairn. Figure 27. Hemitaurichthys multispinosus, BPBM 13327, 148 mm SL, Pitcairn, 43 m. Figure 28. Centropyge hotumatua, transforming (upper) and juvenile (lower) BPBM 13325, 22.6 and 27.5 mm SL, Pitcairn. Figure 29. Plectroglyphidodon leucozona, BPBM 17052, 86 mm SL, Pitcairn tidepool. aor %. 6 Ye 2 cr . p Pp 8 1. Pp aim 7 1 / M. Vy BPB 32 Figure 33. Cirripectes quagga, BPBM 17070, 41 mm SL, Henderson Island. Figure 34. Entomacrodus niuafoouensis, BPBM 13745, 80 mm SL, Henderson Island. Figure 35. Entomacrodus rofeni, BPBM 13750, 50 mm SL, Ducie Atoll. Figure 36. Praealticus caestus, BPBM 16800, 56 mm SL, pool at Down St. Paul, Pitcairn. Figure 37. Rhabdoblennius ellipes, BPBM 16903, 39 mm SL, pool at Down St. Paul, Pitcairn. Figure 38. Acanthurus nubilus, BPBM 16710, 189 mm SL, Pitcairn, 37 m. ATOLL RESEARCH BULLETIN NO. 462 FISH NAMES IN LANGUAGES OF TONGA AND FIJI BY R. CHRISTOPHER MORGAN ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. AUGUST 1999 vite Le call ee Mel Jere ; “ADRORN TI — d YOTOAIABAY woe, FAIA FISH NAMES IN LANGUAGES OF TONGA AND FIJI BY R. CHRISTOPHER MORGAN! ABSTRACT This article provides a record of fish names from two locations in the central Pacific, these being the Vava’u Islands of Tonga, and Taveuni area in northern Fiji. The Handbook for the Collection of Fish Names in Pacific Languages by G.L. Barnett provided a method for the collection of these data during field work at the two places which culturally are part of West Polynesia. Interviews and discussions with fishers yielded a record of the Tongan language names and Fijian language names of approximately 50 fishes that occur in the waters of these islands. These terms are collated with the English names and scientific identifications. The names are the contemporary ones used by Tongans and Fijians in the 1980s and 1990s. The word list is a reference document for marine scientists, fishers, environmentalists, and other work focussed on the maritime cultures of the Pacific, local languages and the nomenclature of fishes. INTRODUCTION As shown by a current global survey of information on characteristics of marine species with medicinal and tonic food value (Perry 1998, Perry and Vincent 1998), there is substantial interest and significance today in the taxonomy, distribution and identification of reef and ocean species. The marine fauna of the Pacific no doubt will be exploited with greater intensity in the future. Therefore information on known fish species recorded in the field may have value for research on species diversity and conservation. Knowledge and nomenclature of species used by indigenous people is a fundamental component of environmental information. The people of the islands of Tonga and Fiji in West Polynesia exploit reef and offshore zones for many species that have subsistence, ceremonial and commercial value. This paper provides an original record of fish names in the Tongan and Fijian languages that I compiled through consultation with local fisherfolk during field work at these two places. METHOD AND CONTEXT In the past this kind of information has been used by linguists to reconstruct the relations among Pacific Islands’ languages and as a key to understanding ancient settlement and migration. The Handbook for the Collection of Fish Names in Pacific 1 Department of Pacific and Asian Studies, University of Victoria, Victoria, British Columbia, Canada Manuscript received 20 January 1999; revised 10 August 1999 2 Languages by Gary L. Barnett (1978) was designed for this purpose. The old indigenous terms were important in that work and, as C. Clerk (1980) pointed out in his review of the Handbook, several problems arise in trying to apprehend the ancient indigenous terms in the late twentieth century. Contemporary concerns for the marine environment, maintainance of the diversity of species, and potential uses of marine species as foods, tonics and medicines give these data new value and historical relevance, especially since these interests are growing at the same time that foreign names are replacing old terms in many local native languages. The approach used to record these data on fish was to consult men who fished as their main economic activity. The Tongan source was a fisherman aged 36 years and living on Ovaka in 1982, who stated that he learned the names as a youth from his father in Vava'u. The source for the Fiji data also was an active diver and fisherman, aged approximately 30 years in 1996, who reported that he learned the terms on Taveuni fishing with his classificatory brothers, one of whom we consulted on several Fijian terms. In the evenings, after fishing on the reef or processing sea cucumber on shore, we sat together and examined the Handbook illustrations and descriptions as the basis to make a written record of the names of the fishes. A limit on this method is that identification of the fish rely on visual and descriptive recognition of the the fish types represented in the reference material by the indigenous sources. On this basis, the Handbook provided a useful tool for collecting contemporary fish names, and of the 50 fishes listed and illustrated, it was possible to record 48 names in the Fijian language and 49 names in the Tongan language. These consultations with local sources in Tonga and Fiji produced the record of data listed in the accompanying table. The record consists of English names, taxonomic identifications, and the terms given in Tonga and in Fiji. A description of the two sites in the Pacific Islands and orientation to the orthography and pronunciation of the respective languages places the data into a context and assists readers to use these terms for fishes. THE SITES The data are from two specific locations in Tonga and Fiji. The Tonga site was Ovaka Island, in the Vava’u Group of Islands. The Fiji site was Taveuni Island and offshore Nukusemanu Islet in the northeast of the Fiji archipelago. These two sites are approximately 600 miles apart under prevailing southeasterly winds and in ancient times the Tongans and Fijians were in regular contact by canoe voyages. Economic links between Tonga and Fiji related to differences between the environments of the island groups. The Vava’u Group are dry islands that lack rivers and irrigation, while Taveuni is a wet island with many streams and cultivations in the valleys along the shores. The Vava’u Islands are located at 18°40’ S and 174° W. The group is a raised coral formation that evidences a complex geohistory of uplifting and subsidences that 3 created the islands.2 The whole group is tilted from high cliffs on the north shore of the main island ‘Uta Vava’u. From these cliffs, the landscape generally grades down to small coral islands in the south and west and extensive reefs across the fringe of the group. Ovaka, where I collected the Tonga language terms, is one of the small southern islands of this Motu district and its inhabitants have made intensive use of their reefs and waters in collecting and fishing for many purposes, including for subsistence foods, ceremonial foods, and for commodity sale. Women concentrate on the inshore zone along shelfs and shallow reefs. Men undertake fishing by line, net and freedive spearing methods offshore and in the deep lagoon inside the fringing reef islands. By contrast, Taveuni Island, at 16°40’ S_ and 180° longitude, was volcanic in origin. A central ridge of crags and old volcanic craters rises abruptly from the northeastern shore. Many rivers flow from these heights, through narrow estuaries across the shore zone into the sea, and their fresh waters reduce the salinity along the inshore zone. In Fiji there is greater local diversity in water environment conditions than in Tonga. Offshore from volcanic Taveuni, some small dry coral islands rise out of shallow reefs that are rich in marifauna. Sea clans of Fijians rely on these species for subsistence and ceremonial usages and also have come to concentrate on the collection and processing of sea cucumber, a historically important commodity, that has been the object of renewed demand in the 1990s for export to Asia and North America. The gender- spatial division of labour in fishing communities is similar to that described for Tonga. Both the Taveuni, Fiji, and Vava'u, Tonga, sites lie west of the Andesite Line that runs between Fiji-Tonga and Samoa. There are differences in the diversity of animals and birds on Taveuni and Vava’u respectively, with Taveuni having many more species, while the marine fauna represented in the named fishes were reported by local sources at both sites. ORTHOGRAPHY AND PRONUNCIATION Tongan is a relatively homogenous language spoken throughout the kingdom.4 Variations in pitch and tone may be heard in different parts of the country. Comparison of these terms from the Motu district of the the Vava’u Group with names from other parts of Tonga remains to be done. Key points of pronunciation in Tongan relating to consonants are that: the ng is rolled together as in “sing”; the p represents the sound between b and p; also the ¢ stands in the absence of a d, there being 12 consonant sounds represented in the Tongan writing system. Vowel length is marked in Tongan by the use of a macron (~ ) to signify a long vowel. 2This description is based on my study and observations on Vava'u. For a discussion of the geology of the Tonga Islands see Ewart and Bryan (1973). 30n the geology of Taveuni Island, see Latham and Denis (1980) and T. Bayliss-Smith et al (1988:13-43). 4 For discussion of features of the Tongan language.and writing system, see the authoritative dictionary of Churchward (1959). Table 1: Names of Fishes in Tonga and Fiji No English 1. ‘tiger shark 2. grey reef shark 3. hammerhead shark 4. eagle ray 5. manta ray 6. moray eel 7. emperor 8. emperor 9. surgeonfish 10. striped surgeonfish 11. unicornfish 12. rabbitfish 13. rudderfish 14. angelfish 15. squirrelfish 16. batfish 17. lunar tail cod 18. blue spotted grouper 19. coral trout 20. giant grouper 21. spotted grouper 22. snapper 23. red snapper 24. oilfish 25. jobfish Taxonomic Galeocerdo cuvieri Carcharhinus amblyrhynchos Sphyrna lewini Aetobatus narinari Manta birostris Gymnothorax sp. Lethrinus olivaceous Lethrinus erythracanthus Ctenochaetus striatus Acanthurus lineatus Naso unicornis Siganus argenteus Kyphosus cinerascens Pygoplites diacanthus Sargocentron sp. Platax teira Variola louti Cephalopholis argus Cephalopholis miniatus Epinephelus lanceolatus Epinephelus tauvina Lutjanus sp. Lutjanus sp. Ruvettus pretiosus Aprion virescens sifisifi naivatu ngatala kula ngatala ‘uli ngatala pulepule beleni dawa mulu sirisiriwai sigeleti corocoro draunavonu tinani drala Table 1: Names of Fishes in Tonga and Fiji (continued) No. English 26. rainbow runner 27. dolphinfish 28. great barracuda 29. striped marlin 30. sailfish 31. bluefin trevally 32. trevally 33. silver scad 34. skipjack tuna 35. yellow-fin tuna 36. flying fish 37. needlefish 38. porcupine fish 39. pufferfish 40. goatfish 41. milkfish 42. grey mullet 43. maori wrasse 44. filefish 45. triggerfish 46. parrotfish 47. parrotfish 48. parrotfish 49. anenomefish 50. remora Source: RCM. Vava’u, Tonga Field Records, 1982; Taveuni, Fiji Field Records, 1996. Taxonomic Elagatis bipinnulata Coryphaena hippurus Sphyraena barracuda Tetrapturus audax Istiophorus platypterus Caranx melampygus Caranx sexfasciatus Selar crumenophthalmus Katsuwonus pelamis Thunnus albacares Cypselurus naresii Strongylura leiura Diodon hystrix Arothron hispidus Parupeneus sp. Chanos chanos Mugil cephalus Cheilinus undulatus Aluterus scriptus Balistoides viridescens Scarus rubroviolaceus Chlorurus sp. Chlorurus microrhinos Amphiprion chrysopterus Remora remora teliteli' uli drodrolagi Sa’u vorowaqa sa’u laca saqa dina saqa drau tugadra ia seu 6 Fiji contains an east - west division of two languages, and dialectal variations.> The national dialect of Fijian is built on the Bau dialect. The following points are general conventions in the prounciation and writing of Fijian. The b is pronounced mb as in “number”; the c represents the th sound as in”’that”; the n represents nd sound as in “end”; the g stands for the ng sound as in “sing”; and the q is pronounced ng as in “finger”. In the Wainikeli district of northern Taveuni Island in Fiji, people who have attended school generally speak the national dialect of Fijian. Adults know and use a local dialect which is a little different from national Fijian, mainly in a few regular consonant shifts including the use of a glottal stop (’ ) in place of the k, that occurs in the national form of Fijian.6 Dixon’s (1988) study considers the dialect of nearby Bouma in relation to the national Fijian language. The Fijian names for the fishes are recorded in the local form of the language used customarily in the Wainikeli district of Taveuni Island.” NAMES AND IDENTIFICATION The identifications of the fishes listed in the table are based on the taxonomic and photographic information provided in Barnett (1978) and in Randall, Allen and Steene (1990), which proved necessary to resolve certain difficulties with taxonomy and English common names encountered in the Handbook. The identifications in the working record in the field were accurate as far as the illustrations and descriptions in the Handbook accurately indicated the fish shown. Where the species attribution seemed questionable, I have kept identification at the genus level. The English names vary by location and as far as possible the list employs the common names most familiar across the Islands region in reference to these Pacific fishes. Table 1 records first the English names, the taxonomic names in the second column, then the terms provided in the local languages at the sites of Vava’u, Tonga, and Wainikeli, Taveuni Island, Fiji, as described. OBSERVATIONS Tongan fishers gave the term Jolau for the profile of the remora, and laumea for the suckers of the remora viewed from the top. The filefish was not seen in the waters of the Vava'u Islands in Tonga. The Fijian sources could not provide a name for the anenomefish though said it is possible some old men might have a name. In Vava’u, Tonga, they recognised the anemonefish but said they did not have a name nor catch the anenomefish as it is too small. 5On these features of pronunciation of the local Fijian languages, the interested reader would best consult a native speaker of the language and also examine the official dictionary of Capell (1973) and the word list of Dixon (1988). 6Dixon (1988) and Geherty (1983) provide in-depth studies of Fiji area dialects. Tin my composition, I write Waini’eli, with a glottal stop, to refer to the language unit and Wainikeli, with ak, to refer to the district. The offshore islet Nukusemanu is written on the charts in the national dialect of Fijian and can shift to Nu’usemanu in the customary dialects of north Taveuni. The name Nu’usemanu means “Isle of Birds” as manu denotes animals inclusive of birds and the islet is a nesting place for many species of marine birds including frigate birds, boobies and albatrosses. 7 Correspondences between the Tongan and the Fijian fish names recorded are limited. There are several recogniseable cognates, including for example mahimahi and maimai for the dolphinfish, haku and sa’u for the needlefish, kanahe and anace for the grey mullet, hwnu and cumu for the triggerfish, etc. Such correspondences as are evident between the Tongan and the Fijian terms do not appear to be particularly systematic for any groupings of the species, morphology or behaviour, beyond the point that fai and the cognate vai are generic terms for the rays in Tongan and Fijian languages respectively. Regularity between the two languages lies in the predictable consonant shifts in those recogniseable terms that are shared. Fiji shows some intrusion from English, as in "tuna" for yellowfin and that the sources were uncertain about i’a seu for skipjack reporting that today Fijians generally call it tuna as well. These intrusions reflect the commercial importance of these fish in Fiji. Indeed, the term tuna is used loosely in Fiji today to refer to at least two species, a trend that highlights the historical value of lists of local names. It is known from other word lists based on field work in Polynesia that local Polynesian sources may insert terms, often in humour, during vocabulary work. A cited example is the experience of Labillardiere, a "natural philosopher" of the French Enlightenment, who recorded terms for Tongan numerals, and communicated his findings to the Academy of Sciences in Paris, not knowing that the terms for the higher numbers included severai obscene words in the Tongan language (see Freeman 1992: 6). While I cross-checked these terms as far as possible in the field to confirm their validity, the fish names presented in this report, as with all data, are subject to verification. CONCLUSION Interest in the diversity and sustainability of species is growing worldwide. As these concerns develop further, the living languages and knowledge of the indigenous people of the Pacific Islands have an importance for general and scientific knowledge. The short survey provides a record of the names of these types of fish as reported by indigenous people in Tonga and Fiji. New names are being incorporated into the local languages. Therefore the record may be of historical relevance. The data have been set in their area and language context to assist study. ACKNOWLEDGEMENTS A special appreciation extends to the sources in Tonga and Fiji, in particular Toni Tupouto'a on Ovaka, Vava'u, and Pio Rova of Pagai in Wainikeli, Taveuni, for their work and cooperation in making this record. Also, I should like to acknowledge the comments of the reviewers that assisted the revision of this interdisciplinary paper. The field studies in Tonga were supported by the Research School of Pacific Studies of the Australian National University and field work in Fiji by the Centre for Asia-Pacific Initiatives of the University of Victoria, Canada. REFERENCES Barnett, G. L. 1978. Handbook for the Collection of Fish Names in Pacific Languages. Pacific Linguistics Series D. No. 14. Research School of Pacific Studies, The Australian National University. Bayliss-Smith, T. et al 1988. Islands, Islanders and the World: The Colonial and Post- colonial Experience of Eastern Fiji. Cambridge, New York, Melbourne: Cambridge University Press. Capell, A. 1973. A New Fijian Dictionary. Suva: Government Press. Churchward, C. M. 1959. Dictionary. Tongan - English, English - Tongan. Nuku’alofa: Government Press. Clerk, C. 1980. Review of Handbook for the Collection of Fish Names in Pacific Languages, by G. L. Barnett. In Journal of the Polynesian Society 89: 380-382. Dixon, R.M.W. 1988. A Grammar of Bouma Fijian. Chicago and London: The University of Chicago Press. Ewart, A. and W. B. Bryan 1973. The Petrology and Geochemistry of the Tonga Islands. In The Western Pacific: Island Arcs, Marginal Seas, Geochemistry. Nedlands, W.A.: University of Western Australia Press. Freeman, D. 1992. Paradigms in Collision. Canberra: Research School of Pacific Studies, The Australian National University. Geherty, P. A. 1983. A History of the Fijian Languages. Oceanic Linguistics Special Publication No. 19. Honolulu: University of Hawaii Press. Latham, M. and B. Denis 1980. The Study of Land Potential: An Open-Ended Enquiry. In Population - Environment Relations in Tropical Islands: The Case of Eastern Fiji. Edited by H.C. Brookfield. MAB Technical Notes 13. Paris: Unesco. Perry, A. 1998. Global Survey of Marine and Estuarine Species Used for Traditional Medicines and/or Tonic Foods. Traditional Marine Resource Management and Knowledge. Information Bulletin 9. Noumea: The South Pacific Commission. Perry, A. and A. Vincent 1998. Survey Re: Marine Species Used for Traditional and Tonic Medicines. Department of Biology, McGill University, Montreal, PQ, Canada. Randall, J.E., G. R. Allen, and R.C. Steene 1990. Fishes of the Great Barrier Reef and Coral Sea. Honolulu: University of Hawaii Press. ATOLL RESEARCH BULLETIN NO. 463 THE NON-NATIVE VASCULAR PLANTS OF HENDERSON ISLAND, SOUTH CENTRAL PACIFIC OCEAN BY STEVE WALDREN, MARSHALL I. WEISLER, JON G. HATHER AND DYLAN MORROW ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. AUGUST 1999 Fora WIE c uosiepueH oue 2 : dnol5 seiqwes dnoryy uredjyig aera A . - dnoip wieoyig’ ** - ‘| owen “sar, « ve ‘| sesenbiew ‘%*. ‘PsoBedejey "| Nei00s UkIIO IIFIIeg YNOS THE NON-NATIVE VASCULAR PLANTS OF HENDERSON ISLAND, SOUTH-CENTRAL PACIFIC OCEAN BY STEVE WALDREN!, MARSHALL I. WEISLER?, JON G. HATHER? AND DYLAN MORROW‘4 ABSTRACT Henderson island, a World Heritage Site in the Pitcairn group, south-central Pacific Ocean, has often been thought to have a pristine vegetation. Our archaeological investigations and field observations in 1991-1992, supported by recent observations in 1997, suggest the occurrence of former areas of Polynesian cultivation near to the North and East Beaches, and indicate that about 17 non-native vascular plant taxa have occurred. The majority of these were deliberate Polynesian introductions, some taxa are known only as sub-fossils from Polynesian occupation sites; some of this sub-fossil material may represent imported plant parts such as timber or food wraps, rather than indicating in situ cultivation. These Polynesian introductions show little spread from their probable site of introduction and are mostly restricted to the vicinity of the northern beaches; some have become extinct on Henderson. The Pitcairn islanders have also introduced a small number of vascular plants, and one of these (Passiflora maliformis) is potentially invasive. Other taxa have been accidentally introduced, at least one by a recent scientific expedition in 1991. The intact nature of much of the native vegetation may have restricted the opportunities for more widespread colonization; care is needed to ensure that this Situation persists. INTRODUCTION Henderson island (24°20’ S, 123°20’ W; 37 km2) is one of four widely-separated islands of the Pitcairn group, lying east of the Tuamotu archipelago, near the southeastern margin of Polynesia. The island is a raised atoll (or makatea island) with the fossil reef surface now some 30 m above Sea level, and most of the coastline is of sheer cliffs, often ! Trinity College Botanic Garden, Palmerston Park, Dartry, Dublin 6, Ireland. Email swaldren@tcd.ie 2 Department of Anthropology, University of Otago, P. O. Box 56, Dunedin, New Zealand. Email marshall.weisler@ stonebow.otago.ac.nz 3 University College London, Institute of Archaeology, 31-34 Gordon Square, London WC1H OPY. Email tcfa289 @ pop-server.bcc.ac.uk 4 Botanic Garden & Department of Pharmacognosy, Trinity College, Dublin 2, Ireland. Email dmorrow @tcd.ie Manuscript received 16 July 1998; revised 5 January 1999 2 considerably undercut. The narrow beaches in the northern part of the island (Fig. 1) provide the only practical means of access from the sea. Apart from the beaches, Henderson is continuously vegetated, except for some bare patches of coral rubble in the former central lagoon patch reefs, and the areas of limestone pinnacles in the northwest (Waldren, Florence & Chepstow-Lusty, 1995b). There is no standing fresh water, although drips are found in the backs of some caves which last for several days after hard rains. The climate is essentially subtropical; data collected in 1991-1992 suggest a seasonal temperature range from a maximum of 31.4 °C (February, March) to a minimum of 12.0 °C (September), wind was predominantly from the east, but wind speed, direction and rainfall varied unpredictably throughout the recording period. For further climatic details see Spencer (1995). Despite Henderson's hostile terrain and limited fresh water, the island's remoteness is probably the key factor that helps to preserve most of the native flora and fauna today. Of the 72 vascular plants recorded, 9 are considered endemic (Waldren, Florence & Chepstow-Lusty, 1995a). There are four endemic birds in addition to breeding populations of various seabirds, some of which are very locally distributed (Brooke, 1995). At least 300 terrestrial arthropods occur, many of which are endemic (Benton, 1995; Benton & Lehtinen, 1995), and Preece (1995) recorded at least 16 terrestrial molluscs, over half of which were considered endemic. These local or endemic taxa, the intact vegetation cover compared to other Pacific raised atolls, and the biogeographic importance of the island’s position near the eastern extremity of the Indo West-Pacific biotic province (Kay, 1984) all contribute to the conservation value of the island, which has been designated a World Hentage Site. The supposed pristine nature of the flora and fauna have attracted much attention (Diamond, 1995; Fosberg, Sachet & Stoddart, 1983). However, the North and East Beaches in particular are generally backed by cliffs with numerous caves and these have been utilised as occupation sites by Polynesian settlers, either transiently or continuously, between ca. AD 1050 and 1600 (Weisler, 1994, 1995, 1996, 1997). A major habitation midden was identified along the North Beach and measures at least 10,000 m2 (Weisler, 1995, 1998). Although these caves and beach areas provided habitation loci for Polynesian groups, they explored much of the island as evidenced by a rock shelter at the remote southern end of the island (Weisler, 1995). Prehistoric gardening areas near the cliff margins on the north and east coasts, and much less on the northwest side of the island, bear evidence of shifting cultivation: stone clearance mounds and charcoal particles dispersed throughout subsoil dated, in some places, to the 13th century (Weisler, 1995). Localised burning for land clearance may have reduced habitat for land and sea birds (Wragg & Weisler, 1994), although direct predation (Weisler & Gargett, 1993) and the introduction of the Pacific rat (Rattus exulans) also contributed towards the extinction of certain bird taxa (Wragg & Weisler, 1994; Wragg, 1995). Alteration of pristine islands throughout the course of human colonization and settlement is well documented (for recent summaries, see papers in Kirch & Hunt, 1997). Consistent with human colonization of Oceanic islands, Polynesian settlers of Henderson introduced new plant species for construction, medicinal and spiritual uses in addition to food. Along with the oceanic staples such as coconut (Cocos nucifera), both candlenut (Aleurites moluccana) and ti (Cordyline fruticosa) were particularly well documented in prehistoric cultural deposits on Henderson. The Pitcairn islanders first visited Henderson in 1843, and continue to regularly collect the wood of Thespesia populnea and Cordia subcordata, used in carving curios. 3 The islanders have planted coconut in several places, especially at the west end of the North Beach, and have attempted to introduce a small number of other species, such as Musa sp. Other temporary visitors to the island include shipwrecked sailors (notably the crew of the Essex; see Fosberg et al., 1983), visitors from passing yachts and pleasure cruises, and recent scientific expeditions. All of these, together with former Polynesian occupation, have affected the vegetation, and may have either deliberately or accidentally introduced plant species. The Sir Peter Scott Commemorative Expedition to the Pitcairn Islands (PISE) visited Henderson for fifteen months during 1991-1992 with several objectives, including biological, archaeological and geological surveys (Benton & Spencer, 1995). A short visit to Henderson was also possible during the 1997 Botanical and Entomological Expedition to Pitcairn. The greater knowledge now available permits a more detailed understanding of the island’s biota and ecology. The aim of this paper is to document the plant taxa introduced prehistorically by Polynesians and other plants species brought to the island after first European contact in 1606 (Quiros, 1904). It is often difficult to ascertain whether a taxon is native or introduced, and typically circumstantial evidence from the local distribution, dispersal mechanisms, potential use and biogeography must be balanced. We therefore provide notes on plant species, together with mapped distributions of the taxa; the latter may assist future scientific visitors to the island in updating information on the status of these taxa. EVIDENCE FOR POLYNESIAN CULTIVATION AREAS The archaeological evidence has been summarised by Weisler (1995). Two sites were located above the North Beach on the plateau, and two sites on the plateau above the East Beach (Figure 1). Local abundance of Cordyline fruticosa at these sites provides a useful marker, and may be a relict of cultivation. Surface gardening indications also included stone clearance mounds. Weisler excavated areas adjacent to these mounds and near the Cordyline stands revealing charcoal flecks dispersed through the subsurface sediments, which is typical of prehistoric gardening areas found elsewhere in the Pacific. The vegetation of the former gardening areas above the East Beach consists mostly of ferns, low shrubs and the weedy native Senecio stokesii. This suggests that once cleared of the typical forest dominated by Pisonia grandis (Waldren et al., 1995b), full forest regeneration has not occurred after abandonment of the plots, possibly due to the harsh conditions, especially the onshore trade winds which buffet the East Beach and plateau. The general topography at these East Beach sites is a gentle slope towards the plateau margin and the sea; above the North Beach the plateau surface is more level, which has possibly favoured better forest regeneration by reducing exposure. LIST OF NON-NATIVE TAXA Achyranthes aspera var. pubescens (Mogq.) Townsend (Amaranthaceae) This taxon has only been recorded from a very limited part of the plateau forest about 1 km south of the North Beach (Fig. 2a). It is possibly a Polynesian introduction, but as it is absent from many areas of known Polynesian activity, it may be native or at least a casual introduction. The fruit is strongly reflexed at maturity and readily attaches to clothing or birds feathers (Fosberg & Renvoize, 1980). A. aspera var. aspera has been 4 recorded on nearby Pitcairn (St. John, 1987); in 1991 it was recorded at a single site and presumed a Polynesian introduction (Florence, Waldren & Chepstow-Lusty, 1995) although it could not be found on Pitcairn in 1997 (S. Waldren & N. Kingston, unpublished data). A similar uncertainty holds for Tonga where the plant has been used medicinally against infections (Whistler, 1991a), it was also used medicinally in Samoa and Niue (Uhe, 1974; Whistler, 1984; Géthesson, 1997). Given the relatively few references to this species in Polynesian ethnobotany, it seems likely that it may have been introduced passively by Polynesian visitors or more probably by birds to Henderson; St. John & Philipson (1962) consider A. aspera an accidental introduction. Aleurites moluccana (L.) Willd. (Euphorbiaceae) This species was collected by the Whitney expedition in 1922, but has not been seen since despite extensive collecting by St. John and Fosberg in 1934 during the Mangarevan Expedition (St. John & Philipson, 1962), or during the PISE (Florence et al., 1995). However, during a brief visit to Henderson in 1997 fragments of periderm (‘shell’) were found on the strand line of North Beach by J. Starmer and S. Waldren; searches of the North Beach embayment forests again failed to locate a living specimen of this obvious species. A. moluccana is native to S.E. Asia and aboriginally introduced throughout Polynesia (Whistler, 1991b). Handy & Handy (1972) claim that the large nut does not float and is not ‘resistant’ to sea or fresh water. It was probably introduced to Henderson by Polynesians as it had many uses: the seed has a high oil content, and was used widely as a source of oil and burnt on skewers for light (Handy & Handy, 1972). The soot from burnt nuts was an important dye (Handy & Handy, 1972; Whistler, 1991a,b), the inner bark juice yielded a reddish-brown dye (Gothesson, 1997), and the plant had various uses in Polynesian medicine (Barrau, 1961; Cox, 1991; Géthesson, 1997). Shell fragments were found in cultural deposits in rock shelters on the North and East Beach (Fig. 2b) dating from as early as AD 1290 to 1440 (Weisler, 1995: Table 2: Beta-45598). Candlenuts were also burnt on Pitcairn island as a source of light (R. Warten, pers. comm.), and it is possible that the tree(s) found in 1922 were introduced by early visitors from Pitcairn, as claimed by St. John & Philipson (1962). Barringtonia asiatica (L.) Kurz (Leycithidaceae) B. asiatica is a widespread strand plant which is probably a Polynesian introduction to the Pitcairn group. It still occurs, albeit in small numbers, in the vicinity of Adamstown on Pitcairn, but living plants have never been recorded from Henderson. The fibrous fruits readily float in seawater and are commonly found on drift lines, however Whistler (1991) suggests it may be a Polynesian introduction to atolls although it iS native on rocky coasts as far east as the Marquesas. Crushed or grated fruits were widely used as a fish poison (Brown, 1935; Whistler, 1991b), and the timber used for light construction (Banack & Cox, 1987), canoes (Brown, 1935) and also for firewood. The fruits were also used as floats for fishing nets (Brown, 1935). Wood specimens were identified from an East Beach rock shelter (Fig. 2c) with associated dates between AD 1330 to 1650 (Weisler 1995: Table 2: Beta-45600). Cocos nucifera L. (Arecaceae) Coconuts occur in well established groves on the North and Northwest beaches (Fig. 2d). There is a small grove on top of the plateau at the North Beach directly opposite the reef pass, and a few small individuals scattered elsewhere about the North Beach plateau margin. There are a few recent plantings on the East Beach, and a few isolated individuals occur in the plateau forest, one juvenile specimen was found 4 km from the North Beach and may represent an ecologically unsound marker for the end of someone’s trail (Operation Raleigh?): these were destroyed when found in 1991-1992. In Se 5 addition, fruits of Cocos and Thespesia were scattered from a helicopter over the northern part of the plateau forest in 1966 during an American airfield survey (M. Fraser, pers. comm.); fortunately few if any seem to have survived. C. nucifera is probably native to the Old World tropics, but has been extensively naturalised throughout the tropics, often leading to secondary colonization or local spread. The species was aboriginally introduced to Polynesia, where it remains a plant of major economic importance (Whistler, 1991a). Although Kirch (1991) lists it as a non-staple throughout Polynesia, coconut was undoubtedly a major staple food on Polynesian and Micronesian atolls, though of less importance on high islands (Barrau, 1961). It has a wide variety of other uses, including beverage, cordage, matting, thatch and timber (Kirch & Yen, 1982; Whistler, 1991a). It was most likely brought to Henderson by Polynesian settlers as the shell, wood, and husk of the plant was identified at several sites on the North and East beaches. The earliest associated radiocarbon dates are calibrated between AD 1000 to 1390 (Weisler, 1995: Table 2: Beta-59983). However, plants introduced by Polynesians may well have died out or at least become very rare. Captain Beechey, who visited in the Blossom in 1825, did not mention this obvious tree but stated that the tallest trees, and the only ones to yield edible fruit, were Pandanus (Fosberg et al., 1983). Coconuts were introduced during the first visit by the Pitcairn islanders in 1843 (Fosberg et al., 1983), the current groves of coconut on Henderson probably derive entirely from plantings made by Pitcairn islanders supplemented with some local dispersal. Cordia subcordata Lam. (Boraginaceae) On Henderson this species is restricted to forests occupying the sandy areas behind the dunes of the North and East Beaches (Fig. 2e); Cordia is apparently absent from similar habitats at the Northwest Beach. A widespread Indo-Pacific strand plant, the seeds are dispersed by flotation (Fosberg & Renvoize, 1980), and so the species might possibly be native. However, Whistler (1991b) claims it is an aboriginal introduction in the eastern part of its range, including Hawai‘i. The nuts are eaten in the Tuamotu islands and elsewhere, mostly in times of famine (Barrau, 1961; Whistler, 1992), and the leaves provide occasional food on Tikopia (Kirch & Yen, 1982), but the main value of the plant is for its beautiful finely grained timber which is highly valued for construction of houses, boat parts and for various artefacts and handicrafts (Whistler, 1991a, 1992). The inner bark fibres have been used in weaving, and various parts of the plant have been used medicinally (Whistler, 1992). The attractive orange flowers may also have been used by Polynesians for making /ei. Pitcairn islanders have regularly harvested Cordia from Henderson for use in their wood carving industry; the few large trees seen on the North Beach in 1991 had all been removed by 1997, although smaller trees still occur on the cliff slopes of the North and East beaches. Probably a Polynesian introduction, but see also comments on Thespesia populnea below, which occurs in similar habitats. Cordyline fruticosa (L.) Chev. (Agavaceae) Cordyline is largely restricted to the plateau margin (Fig. 2f). It occurs at the top of the cliff directly above most caves which have prehistoric Polynesian occupation on the North Beach. It is less frequent at the East Beach, and occurs several hundred meters inland from the cliff edge, possibly marking former cultivation areas. There are isolated groups of plants on the plateau margin along the east coast, and at the northwest point. There is also one large grove of individuals at the North Beach just below the cliff at the site of the PISE camp, this location was the earliest habitation site on Henderson (Weisler, 1995). Leaves and wood of Cordyline were found in several rock shelters on 6 the North and East beaches with earliest associated dates between AD 1330 to 1650 (Weisler, 1995: Table 2: Beta-45600). C. fruticosa is probably native to tropical Asia and is an aboriginal introduction to Polynesia (Whistler, 1991b). Handy & Handy (1972) claim it to be native to Hawai‘1, but this is highly unlikely and the species was treated as a Polynesian introduction by Wagner, Herbst & Sohmer (1990). It is widespread throughout Polynesia, and many cultivars are known, for example those with red leaves which may be grown purely for omament. The roots were widely used as a sugar source (Whistler, 1991a), containing ~20% sucrose according to Barrau (1961), who also states that cooked roots keep well. The leaves were used as food wraps, cordage and clothing, and the plant was used in herbal medicine (Handy & Handy, 1972; Kirch & Yen, 1982; Whistler, 1991a; Whistler, 1991b), but Cordyline also had great religious significance (Whistler, 1991b). For example, in Hawai‘i the plant was used for altar decoration, protection from evil spirits and for exorcism (Handy & Handy, 1972), and perhaps of more significance to Henderson, was used as a boundary marker in Tikopia (Kirch & Yen, 1982). Cordyline observed in 1997 growing high on the cliffs at Tautama, the major stone quarrying area of Pitcairn, may well have been planted for similar marker/religious purposes. The distribution of Cordyline on Henderson suggests it was used as a food supplement and is often found at gardening locations (Weisler et al., 1991), and the plant probably had great religious and cultural significance. No flowering or fruiting has ever been recorded from Henderson Cordyline, which may represent a sterile cultivar resulting from a single Polynesian introduction; nearby Pitcairn has many fully fertile individuals. Cyrtosperma chamissonis (Schott) Merr. (Araceae) Cyrtosperma is known from Henderson as subfossil leaf material from two rock shelters on the North Beach; it is absent from the modern flora and it is possible that it never grew on the island. The Henderson specimens were found in prehistoric cultural deposits associated with earth ovens (Fig. 2g) dated from AD 1330 to 1648 (Weisler, 1995: Table 2: Beta-59009). Indeed, Cyrtosperma leaves were often used to wrap items for transport and to cover earth ovens during cooking. The species is native to New Guinea, but aboriginally introduced to Polynesia (Whistler, 1991b). Aroid tubers remain an important starch source, but Cyrtosperma was of less importance than Colocasia esculenta (Kirch, 1991; Whistler, 1991b)—at least on high volcanic islands. However, its ability to grow in brackish water made it a major staple on atolls (Barrau, 1961), although at least some cultivars can grow readily in drier conditions, as on Tikopia (Kirch & Yen, 1982). The occurrence of leaf material suggests it was cultivated in situ, although leaves may possibly have been imported from Pitcairn or Mangareva. Hedyotis romanzoffiensis (Chamisso. & _ Schliechtendal) Fosberg (Rubiaceae) A specimen of this species with immature fruit was first found in 1997 by S. Waldren, who had previously collected seed from the Oeno island population (St. John & Philipson, 1960) in 1991. The species was found growing in a sandy substrate slightly east of the Pitcairners camp at North Beach, in an area used for drying seed in 1991 (Fig. 2h). As the plant was growing in an appropriate ecological niche, and as Henderson is within the natural range of the species, the individual plant was not removed. A sample of leaf tissue was dried in silica gel for DNA extraction to allow any colonization by the taxon to be followed genetically. Hernandia sp. Hernandia stokesii occurs as a supposedly native species on Henderson, and is restricted to the highly dissected makatea inland from the Northwest Beach (Fig. 2i). It may be conspecific with H. sonora which occurs in many of the valleys on Pitcairn, where its status is uncertain. As with many of the other fossil timbers, it is not clear whether this represents a Polynesian introduction of timber extracted elsewhere, or whether plants were grown on Henderson. The light timber was used for outrigger floats (Banack & Cox, 1987), and was until recently used for making light fishing canoes on Pitcairn (Reynold Warren, pers. comm.); other Hernandia species have been similarly used in the Marquesas (Brown, 1935). Charred wood of a Hernandia species was recovered from an East Beach rock shelter with an associated date of AD 1330 to 1650 (Weisler, 1995: Table 2: Beta-4560). Although it is possible that Polynesians may have caused the local extinction of this species from the vicinity of the East Beach, it seems more likely that the wood was imported from elsewhere, possibly from Pitcairn. It is unlikely that the Polynesians were exploiting the native H. stokesii which has a highly restricted distribution. Hibiscus tiliaceus (Malvaceae) H. tiliaceus is widespread in Polynesian coastal forests, including Pitcairn (Florence et al., 1995; S. Waldren & N. Kingston, unpublished data). The species has never previously been recorded from Henderson; as there have been few archaeological records and no modern records, we consider the Henderson material may be of Polynesian origin. However, Wagner et al. (1990) considered this species as indigenous to the Hawai’ian and Marquesas islands, and it may well have been native to Henderson. Also, it is not clear whether the fossil Henderson material represents imported wood, or whether it was of local origin. Identified from charred wood from an East Beach rock Shelter (Fig. 2) with an associated date of AD 1280 to 1430 (Weisler, 1995: Table 2: Beta-45601). The inner bark fibers were used as cordage (Brown, 1935; Banack, 1991; Whistler, 1991b), and the light timber used for construction of floats for canoes (Banack, 1991) and fishing nets (Whistler, 1992); the species was also used medicinally (Brown, 1935; Cox, 1991; Whistler 1992). The large leaves served as plates or food wraps (Brown, 1935); on Pitcairn they have occasionally been used as a substitute for toilet paper (Meralda Warren, pers. comm). Lycopersicon esculentum Mill. (Solanaceae) A few tomato plants have on occasion been introduced by the Pitcairn islanders; none seems to have survived in the sandy soils of the North Beach area for very long (Fig. 2k). Musa sp. (Musaceae) Bananas have been introduced by the Pitcairners but, like tomato, they do not persist for long. In 1991-1992 one was growing in an oil drum ‘pot’ near the North Beach landing, it was absent in 1997; it is unlikely that the plant would survive to produce fruit on Henderson. The large leaves are used to wrap food for transport, for cooking, and for covering earth ovens, while the fruit provides an important food. Leaves of Musa sp. were identified at an East Beach site (Fig. 21) with an associated date of AD 1410 to 1660 (Weisler, 1995: Table 2: Beta-45602). It is not clear whether the sub-fossil material from Henderson represents food wraps brought from outside; it is likely that Polynesian settlers brought living plants with them, but these failed to persist. Pandanus tectorius Parkinson (Pandanaceae) Pandanus groves occur throughout Henderson (distribution not mapped), from the most hostile locations on the southwestern cliffs, to the shelter of the mixed Pisonia 8 forest of the plateau. St John described many local species of Pandanus, including P. hendersonensis, but these are all currently considered to be part of P. tectorius. There is considerable variation present between the Henderson individuals, although each grove seems to be reasonably uniform. P. tectorius is widespread through the Indo-Pacific region (Fosberg & Renvoize, 1980). The status of Pandanus on Henderson is difficult to ascertain. The fruits are naturally dispersed by sea and the seed may remain viable after flotation in seawater for eight weeks (Fosberg & Renvoize, 1980), so there is likely to be a native population on Henderson. However, the species was: of great importance to Polynesians and had a great many uses. Timber was used for house frames (Whistler, 1991a), although in Hawai‘i male trees are claimed to yield better timber than females (Handy & Handy, 1972). The leaves were used for thatch and, after removing thorns, for fine weaving (Handy & Handy, 1972; Whistler, 1991b), as they still are on Pitcairn island. Pollen was used to scent oil (Whistler 1992), and the leaves (Uhe, 1974) and aerial roots (Handy & Handy, 1972; Whistler, 1991b) were used medicinally. The plant was eaten in various ways in Polynesia (Handy & Handy, 1972; Whistler, 1991a), probably most often as a famine food (Whistler, 1991b). The soft base of the phalange was eaten as was the small seed in each fruit, and the hearts of the terminal branches were eaten after steeping in seawater to remove calcium oxalate (Barrau, 1961). As with Aleurites and Cocos, it is almost certain that Polynesians colonising or even visiting Henderson would have brought Pandanus with them, because of its great value. Introduced Polynesian cultivars may have integrated with the existing native population, resulting in a wide spectrum of morphological variation. Fosberg has suggested that variation in P. tectorius resembles that found in a highly variable horticultural species such as apple (Barrau, 1961). Passiflora maliformis L. (Passifloraceae) Passiflora maliformis was seen at the eastern end of the North Beach in 1991 (Fig. 2k; Florence et al., 1995), and empty fruits found washed up on North Beach in 1997. It is a recent deliberate introduction from Pitcairn island, where the species is invasive (Florence et al., 1995). Various Passiflora species have become invasive weeds throughout the tropics; efforts should be made to eliminate this species from Henderson. Solanum americanum P. Miller (Solanaceae) Solanum americanum was first recorded from Henderson in 1991 from the southern cliffs (Fig. 2n; Florence et al., 1995). As this site is very remote from the access points in the north of the island, it may have been previously overlooked. However, as only a single plant was found, it is likely to be a recent colonist, probably by an avian vector. The species is a widespread weed (Barrau, 1961), and is common in the littoral vegetation of Pitcairn, but it is unlikely to have been introduced by Pitcairn islanders who never visit the southern part of Henderson. A Polynesian rock shelter occurs in the vicinity of the Henderson plant, and as the plant is sometimes used as a vegetable (Barrau, 1961; Whistler, 1991a) and been used medicinally (Géthesson, 1997), it may represent a Polynesian introduction, though we consider this species to be a recent and probably ephemeral colonist. Setaria verticillata (L.) Beauv. (Poaceae) This species was recorded in 1987 and 1991-1992 from the North Beach landing area (Fig. 2n). It is a widespread tropical and subtropical weed, common on nearby Pitcairn from where the Henderson material probably originated. The fruiting spikelets have bristles with reflexed teeth that readily attach to clothing. All plants seen of this recent colonist were destroyed, and it could not be found in 1997. Thespesia populnea (L.) Solander ex Correa (Malvaceae) On Henderson Thespesia is found in the embayment forests of the North and Northwest beaches (Fig. 20) and is an important member of the vegetation there, associated with Pisonia grandis, Celtis pacifica, Glochidion pitcairnense etc; it is apparently absent from the plateau forests (Waldren et al., 1995b). The fruit is an indehiscent capsule; the seeds, and presumably the fruits too, are capable of flotation for two months in seawater (Morrow, 1993), and this is likely to be a major means of dispersal. On Henderson, the seeds and fruit were ignored by rats (Rattus exulans) and species of hermit crab, probably because of toxic substances present in the seed and fruit. The native range of this species is uncertain because of aboriginal introductions; it is thought to be introduced to the eastern parts of Polynesia (Handy & Handy, 1972, Whistler 1991a) and Easter Island (Zizka, 1991). It may have originally been native to S.E. Asia (Zizka, 1991), but is now pantropic (Fosberg & Renvoize, 1980). Thespesia had a variety of uses in Polynesian culture. Bark and leaf extracts were used for treating a variety of ailments (Whistler, 1991b). Trees were often planted around marae (temples) and it may have been important as a sacred tree (Wagner et al., 1990). The species was used in Polynesian herbal medicine (Cox, 1991), mainly used topically or taken internally for intestinal disorders (Morrow, 1997). Various preparations of leaves, bark and heartwood have shown antibacterial, antifungal and antispasmodic pharmacological activity (Morrow, 1997). The wood is hard and durable, and was widely used for construction of durable items, such as bowls and canoe paddles (Whistler, 1991a,b). The yellow flowers are attractive and may have been used by Polynesians for decoration, particularly on Henderson where few species have large or showy flowers. The delicately coloured wood is still used for carving curios in Polynesia, including Easter Island (Zizka, 1991) and Pitcairn. Pitcairners regularly visit Henderson to cut both Thespesia and Cordia, the main timbers used on Pitcairn for carvings. The islanders prefer the Thespesia of Henderson to that of Pitcairn because the heartwood is more intensely coloured. At present it is likely that Thespesia and Cordia are harvested non-sustainably from Henderson. The typical habitat of the species is coastal forests and the species is restricted to this habitat on Henderson, suggesting it may be native. On nearby Pitcairn island Thespesia occurs in a variety of habitats and is much more obviously introduced. Like Cordia and Pandanus, Thespesia may well have been introduced to Henderson by Polynesian settlers, but this does not preclude the existence of native populations. Thuarea involuta (G. Forster) R. Brown This beach grass may be a Polynesian weed, but we consider it more likely to be native, despite sometimes being used ceremonially on Tonga (Whistler, 1992; Géthesson, 1997). It occurs in two distinct habitats: among shrubs along beach fronts at the East and North Beaches, and in coral rubble in fairly open inland Xylosma suaveolens forest south of the North Beach (Fig. 2p). DISCUSSION Of the 17 or so species that definitely or possibly are not native to Henderson island, two are considered to be recent adventives, three or possibly five to be deliberate introductions by the people of Pitcairn island, up to ten are deliberate prehistoric Polynesian introductions, and one is a possible accidental Polynesian introduction. One of the recent adventives (Solanum americanum) probably arrived on the island following 10 avian dispersal, possibly from Pitcairn island where the species is common. If so, this represents natural secondary dispersal of a weedy non-native species. The same may also hold true for Cocos, Cordia, Hibiscus and Thespesia, whose fruits are naturally dispersed by flotation on seawater, but are considered by many to be aboriginal introductions to eastern Polynesia. Three taxa (Barringtonia, Cyrtosperma, and Hibiscus) are known from Henderson only as subfossil material. Randy Christian and Meralda Warren of Pitcairn claim to have found a Piper species on Henderson; this requires further investigation. The identification of subfossil leaves, fruit fragments and wood has added greatly to our understanding of human plant introductions to Henderson and permits a more complete appraisal of the historical relationships between people and plants on isolated makatea islands. There is clear evidence that Polynesian settlers deliberately introduced plants to Henderson island, and they doubtless had very good reasons for doing so. The first settlers would have found few indigenous plants that could be used as food and were probably known to the settlers; these included the leaves of Lepidium bidentatum and Pisonia grandis, and the stem bases of Portulaca lutea which can be eaten as a vegetable (Barrau, 1961), but these are at best occasional or famine foods and not major staples. Even on larger and floristically more diverse islands such as in Tonga, there are few native food plants (Whistler, 1991b). The earliest Polynesian visitors to Henderson may well have been transient harvesters of the previously untapped sea bird colonies and abundant marine resources such as fish, shellfish, and turtles, with permanent settlement occurring at a later time. The need for starch sources probably led to the introduction of Cordyline which also served for clothing and religious purposes. Certain medicinal plants may not have been available locally, and the settlers would doubtless have introduced cultivated forms of species with multiple uses, such as Cocos and Pandanus. As suggested earlier, the introduction of cultigens alongside an existing native population has probably lead to the present range of morphological variation in Pandanus. Probably a variety of cultivars best suited to particular uses were grown. At some period during the Polynesian occupation of Henderson extensive permanent habitations were utilised, and this probably coincided with direct production of crops from gardens. It is clear that the Polynesian settlers introduced a variety of plants which could provide food, timber, clothing, cordage, and medicines; further evidence of carefully planned voyages for settlement. The settlers may have accidentally introduced weed species such as Achyranthes aspera and Thuarea involuta, although the species might equally well have arrived following adherence to birds feathers and the caryopses of the latter are probably capable of flotation. It is difficult to determine whether certain taxa are native or aboriginal introductions. Thespesia populnea and Cordia subcordata may possibly have been present on Henderson prior to initial Polynesian contact, although they are widely considered to be aboriginal introductions in eastern Polynesia (Handy & Handy, 1972; Whistler, 1991b; Whistler, 1991a; Zizka, 1991). Even if these do represent Polynesian introductions, they have integrated with existing beach forest vegetation although they have not spread to the plateau. None of the known or suspected Polynesian introductions are currently invasive, and all are more or less restricted to the northern end of the island, where the highest densities of prehistoric Polynesian habitations are found. This may be because the native vegetation has seen little disturbance over most of the island, thereby restricting the opportunities for successful spread of weedy species. Two other species, considered to be native, are also restricted to the north of the island, these are Caesalpinia bonduc and 11 Senna glanduligera. Caesalpinia has limited Polynesian uses, mainly as a snare for birds or fruit bats (Whistler, 1984, 1991a), while Senna spp. (and the related genus Cassia) are known to have purgative properties. Both species are dispersed by sea, the indehiscent lomentum of Senna floats, and although the very hard seeds of Caesalpinia are not buoyant, the pod probably is and accounts for sea dispersal claimed by Fosberg & Revoize (1980). Both these species are likely to have washed up on the northern beaches, and their subsequent spread south would have been slow as biotic dispersal is likely to be minimal. Polynesian activities do not seem to have had a major effect on the vegetation of the island. Even the known garden areas have become completely invaded by native species, although forest communities have failed to develop on these sites. There is also little evidence to suggest that Polynesian activities caused the local extinction of native plant species, unlike the situation with birds (Wragg & Weisler, 1994; Wragg, 1995); the fossil evidence of Hibiscus tiliaceus in the absence of any living records may be the exception. Polynesian settlers utilised some of the native trees, as evidenced by the presence of charred wood of Nesoluma sp. (Fig. 2m), presumably the endemic WN. st- Johnianum Lam & Meeuse, at the East Beach (Hen 10) with an associated age of AD 1330-1650 (Beta-45600). The more recent effects of excessive cutting of Thespesia by American personnel during an airfield survey have had a much greater impact on the vegetation: areas of beach embayment forest at the North Beach have been colonised by Pandanus, leaving numerous cut Thespesia stumps in the deep leaf litter. Recent introductions by the Pitcairn islanders are restricted to food plants, partly as an emergency in case bad weather necessitates a prolonged stay on Henderson. Some of these introductions are potentially invasive, and may damage the native communities. There are now well established groves of coconut at the North and Northwest beaches, and these seem to be invading coastal forest and dune ridge communities. The presence of Passiflora maliformis and Setaria verticillata indicates the potential problem of weedy species being deliberately or accidentally introduced to Henderson. Both these species are invasive weeds on Pitcairn, and their spread on Henderson should be monitored closely. Thankfully the efforts of the PISE in 1991-1992 seem to have eradicated Setaria, although the botanical phase of the expedition can be rather embarrassingly held responsible for the accidental introduction of Hedyotis from Oeno. Pitcairn has a considerably larger non- native flora than Henderson, most of which are relatively recent introductions, and some are highly invasive. As pointed out previously (Waldren et al., 1995a), great care should be taken to prevent their spread to Henderson. ACKNOWLEDGEMENTS We are grateful to the Pitcairn Island Council and the Pitcairn Island Commissioner for permission to visit the islands and to collect samples and specimens. SW and MW are especially grateful to the Pitcairn Islanders for their generous hospitality during their stay on Pitcairn, and especially thank Reynold & Nola Warren, and Steve & Olive Christian and family. Their visit to the islands as part of the Sir Peter Scott Commemorative Expedition was generously supported by the following major sponsors: The Royal Society, International Council for Bird Preservation, British Ornithologists Union, J.A. Shirley, Foreign & Commonwealth Office UK, UNESCO; other sponsors appear in the expedition report of 1992. MW also thanks the Wenner-Gren Foundation for Anthropological Research. Thanks to Naomi Kingston, John Starmer and Pierre 12 Bingelli for recently searching the Henderson strandline with SW, and thanks to Graham Wragg and Ed Saul of RV Te Manu for logistic support. This is paper no 86 of the Sir Peter Scott Commemorative Expedition to the Pitcairn Islands. REFERENCES Banack SA 1991 Plants and Polynesian voyaging. In: Cox PA and Banack SA, eds, Islands, Plants and Polynesians. An Introduction to Polynesian Ethnobotany, pp.25- 39. Dioscorides Press: Portland, Oregon. Banack SA and Cox PA 1987 Ethnobotany of ocean-going canoes in Lau, Fiji. Economic Botany 41:148-162. Barrau J 1961 Subsistence agriculture in Polynesia and Micronesia. Bernice P. Bishop Museum Bulletin 223:1-94. Benton TG 1995 Biodiversity and biogeography of Henderson Island’s insects. Biological Journal of the Linnean Society, 56: 245-259. Benton TG and Lehtinen PT 1995 Biodiversity and origin of the non-flying terrestrial arthropods of Henderson Island. Biological Journal of the Linnean Society, 56: 262- 272: Benton TG and Spencer T 1995 The Pitcairn Islands: Biogeography, Ecology and Prehistory, 422 + xxxi pp. Academic Press.: London, U.K. Brooke MdeL 1995 The modern avifauna of the Pitcairn Islands. Biological Journal of the Linnean Society, 56: 199-212. Brown FBH 1935 Flora of Southeastern Polynesia. III. Dicotyledons. B.P. Bishop Museum Bulletin, 130: 1-386. Cox PA 1991 Polynesian herbal medicine. In: Cox PA and Banack SA, eds, Jslands, Plants and Polynesians. An Introduction to Polynesian Ethnobotany, pp. 147-168. Dioscorides Press: Portland, Oregon. Diamond J 1995 Introduction to the exploration of the Pitcairn Islands. Biological Journal of the Linnean Society, 56: 1-5. Florence J, Waldren S and Chepstow-Lusty AJ 1995 The flora of the Pitcairn Islands: a review. Biological Journal of the Linnean Society, 56: 79-119. Fosberg FR and Renvoize SA 1980 The Flora of Aldabra and Neighbouring Islands., 358 pp. HMSO: London. Fosberg FR, Sachet M-H and Stoddart DR 1983 Henderson Island (South eastern Polynesia): summary of current knowledge. Atoll Research Bulletin, 272: 1-47. Gothesson, L-A 1997 Plants of the Pitcairn Islands Including Local Names and Uses, 394 pp. Centre for South Pacific Studies, University of New South Wales: Sydney. Handy ESC and Handy EG 1972 Native planters in old Hawaii. Their life, lore and environment. Bernice P. Bishop Museum Bulletin, 233: 1-641. 13 Kay A 1984 Patterns of speciation in the Indo-West Pacific. In: Radovsky FJ, Raven PH and Somer SH, eds, Biogeography of the Tropical Pacific: Proceedings of a Symposium. pp. 15-31. Bernice P. Bishop Museum: Honolulu. Kirch PV 1991 Polynesian agricultural systems. In: Cox PA and Banack SA, eds, Islands, Plants and Polynesians. An Introduction to Polynesian Ethnobotany, 113- 133. Dioscorides Press: Portland, Oregon. Kirch PV and Hunt TL 1997 Historical Ecology in the Pacific Islands: Prehistoric Environmental and Landscape Change, 331 + XV pp. Yale University Press: New Haven. Kirch PV and Yen DE 1982 Tikopia. The prehistory and ecology of a Polynesian outlier. Bernice P. Bishop Museum Bulletin, 238: 1-396. Morrow D 1993 The Biology of Thespesia populnea and its use by the Pitcairn Islanders. Unpublished B.A. (Mod.) thesis, University of Dublin. Morrow D 1997 Evaluation of the ethnomedicinal uses of Thespesia populnea (L.) Sol. ex Correa, using screens to isolate bioactive principles. Unpublished Ph.D. thesis, University of Dublin. Preece RC 1995 Systematic review of the land snails of the Pitcairn Islands. Biological Journal of the Linnean Society, 56: 273-307. Quiros PF de 1904 The Voyages of Pedro Fernandez de Quiros, 1595-1606. 2 volumes, translated by C. Markham. Hakluyt Society: London. Spencer T 1995 The Pitcairn Islands, South Pacific Ocean: plate tectonics and climatic contexts. Biological Journal of the Linnean Society, 56: 13-42. St. John H 1987 An account of the flora of Pitcairn Island with new Pandanus species, 65 pp. Privately published: Honolulu, Hawai‘i. St. John H and Philipson WR 1960 List of the flora of Oeno Atoll, Tuamotu archipelago, South-Central Pacific Ocean. Transactions of the Royal Society of New Zealand, 88: 401-403. St. John H and Philipson WR_ 1962 An account of the flora of Henderson Island, South Pacific Ocean. Transactions of the Royal Society of New Zealand, Botany, 1: 179- 194. UheG 1974 Medicinal plants of Samoa. Journal of Economic Botany, 28: 1-30. Wagner WL, Herbst DR and Sohmer SH 1990 Manual of the flowering plants of Hawai‘i. B.P. Bishop Museum Bulletin, 83. Waldren S, Florence J and Chepstow-Lusty AJ 1995a Rare and endemic vascular plants of the Pitcairn group, south-central Pacific Ocean: a conservation appraisal. Biological Conservation, 74: 83-98. Waldren S, Florence J and Chepstow-Lusty AJ. 1995b A comparison of the vegetation communities from the islands of the Pitcairn Group. Biological Journal of the Linnean Society, 56: 121-141. Weisler M, Benton TG, Brooke M de L, Jones PJ, Spencer T and Wragg G 1991 The Pitcairn Islands Scientific Expedition (1991-1992): first results, future goals. Pacific Science Association Bulletin, 43: 4-8. 14 Weisler MI 1994 The settlement of marginal Polynesia: New evidence from Henderson Island. Journal of Field Archaeology, 21: 83-102. Weisler MI 1995 Henderson Island prehistory: colonization and extinction on a remote Polynesian island. Biological Journal of the Linnean Society, 56: 377-404. Weisler MI 1996 Taking the mystery out of the Polynesian ‘mystery islands’: A case study from Mangareva and the Pitcairn group. In: Davidson JM, Irwin G, Leach BF, Pawley A and Brown D, eds, Oceanic Culture History: Essays in Honour or Roger Green, pp.615-629. New Zealand Journal of Archaeology Special Publication: Dunedin, New Zealand. Weisler MI 1997 Prehistoric long-distance interaction at the margins of Oceania. In: Weisler MI, ed, Prehistoric Long-distance Interaction in Oceania: An Interdisciplinary Approach., pp.149-172. New Zealand Archaeological Association Monograph 21. Weisler MI 1998 Issues in the colonization and settlement of Polynesian islands. In: Vargas Casanova P, ed., Proceedings of the Second International Congress on Easter Island and East Polynesian Prehistory, pp. 79-94. Easter Island Studies Institute, FAU, University of Chile: Santiago. Weisler MI and Gargett RH 1993 Pacific island avian extinctions: the taphonomy of human predation. Archaeology in Oceania, 28: 85-93. Whistler WA 1984 Annotated list of Samoan plant names. Journal of Economic Botany, 38: 464-489. Whistler WA 1991a The ethnobotany of Tonga: the plants, their Tongan names, and their uses. Bishop Museum Bulletin in Botany, 2: 1-155. Whistler WA 1991b Polynesian plant introductions. In: Cox PA and Banack SA, eds, Islands, Plant and Polynesians, An Introduction to Polynesian Ethnobotany, pp. 41- 66. Dioscorides Press: Portland, Oregon. Whistler WA 1992 Flowers of the Pacific Island Seashore, 154 pp. Isle Botanica: Honolulu. Wragg GM_ 1995 The fossil birds of Henderson Island, Pitcairn Group: natural turnover and human impact, a synopsis. Biological Journal of the Linnean Society, 56: 405- 414. Wragg GM and Weisler MI 1994 Extinctions and new records of birds from Henderson island, Pitcairn group, South Pacific Ocean. Notornis, 41: 61-70. ZizkaG 1991 The flowering plants of Easter Island, 108 pp. Palmengarten: Frankfurt, Germany. Table 1. Summary of plant introductions to Henderson island Taxon Achyranthes aspera vat. pubescens Aleurites moluccana Barringtonia asiatica Cocos nucifera Cordia subcordata Cordyline fruticosa Cyrtosperma chamissonis Hedyotis romanzoffiensis Hernandia sp. Hibiscus tiliaceus Lycopersicon esculentum Musa sp. Pandanus tectorius Passiflora maliformis Solanum americanum Setaria verticillata Thespesia populnea Thuarea involuta Likely Origin passive Polynesian introduction or native deliberate Polynesian/Pitcairn introduction Polynesian timber import? deliberate Polynesian/Pitcairn introduction (native?)/deliberate Polynesian introduction deliberate Polynesian introduction deliberate Polynesian introduction accidental expedition introduction native; Polynesian timber import? Polynesian timber import? Native? deliberate Pitcairn introduction deliberate Polynesian and Pitcairn introduction native/deliberate Polynesian introduction recent adventive/deliberate Pitcairn introduction recent adventive recent adventive native/deliberate Polynesian introduction native/Polynesian weed Dispersal Adherence Seawater Seawater Seawater 2 u Flotation or ingestion Avian ingestion? Flotation? ? x Flotation Avian ingestion? Adherence Flotation Flotation or adherence? 24° 20'S Northwest ————————— 1 km (approx) 123° 20' W =) Paths @ Camps South Point tossestes’ Plateau Margin :::| Beaches : Polynesian Cultivation Areas Figure 1. Henderson island, south-central Pacific Ocean, showing places named in text, plateau margin, beaches and areas of Polynesian cultivation. | — Achyranthes aspera : var. pubescens | Aleurites moluccana | d Cocos nucifera | | Figure 2 Barringtonia asiatica Figure 2. Distribution of potentially non-native plant taxa on Henderson. All fossil plant finds (-+), isolated extant individuals or small groups (@) and more continuous distributions (shading) are recorded. Cordia subcordata | Cordyline fruitcosa | Cyrtosperma chamissonis Hedyotis romanzoffiensis Figure 2 Hernandia stokesii Hibiscus tiliaceus Lycopersicum esculentum @ Passiflora maliformis @ Figure 2 Nexbtiateie Setaria verticillata & P- | olanum americanum Thespesia populnea Thuaria involuta Figure 2 ATOLL RESEARCH BULLETIN NO. 464 SOIL-PLANT RELATIONSHIPS AND A REVISED VEGETATION CLASSIFICATION OF TURNEFFE ATOLL, BELIZE BY MALCOLM R. MURRAY, SIMON A. ZISMAN AND CHRISTOPHER D. MINTY ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. AUGUST 1999 MEXICO Caribbean Sea Ambergris HONDURAS |. = - Cay ae, \ GUATEMALA ! * NICARAGUA Turneffe BELIZE ~ pb ow Lighthouse Reef 17°N pes / Gldvers Figure 1 Location map: Turneffe Atoll, Belize Based on maps in Murray (1995) Showing the location of the study site - Turmeffe Atoll, a series of small islands approximately 30 kilometres east of mainland Belize, Central America. SOIL-PLANT RELATIONSHIPS AND A REVISED VEGETATION CLASSIFICATION OF TURNEFFE ATOLL, BELIZE BY Malcolm R. Murray ', Simon A. Zisman ! and Christopher D. Minty ' ABSTRACT This paper presents the findings of an investigation into the land cover and soils of Turneffe Atoll, Belize. Ten vegetation associations and four soil types are identified along with the evolving impact of human activities on the atoll. This information is timely given the increasing threat to this area from new tourist development. An enhanced understanding of Turneffe’s environment enables coastal zone managers to better identify conservation priorities and to actively guide the development taking place. INTRODUCTION Turneffe Atoll lies 18 kilometres beyond the edge of the main Belizean Barrier Reef, which is itself 13 kilometres from the Atlantic coast of mainland Belize, Central America. The least well known of Belize’s three atolls, neither its soils nor its vegetation have been studied in depth. Consequently it remains the most extensive but least studied atoll in the Caribbean. The Turneffe complex is 50 kilometres long and reaches a maximum width of 16 kilometres (Figure 1). It comprises roughly 450 islands, locally known as cays. They range from large complex land masses (the largest island covers 4,493 ha) to small coral rubble cays under 0.02 ha. The islands encircle two large inner lagoons, both of which are generally limited to depths of less than 4 m. Channels (creeks and bogues) link the lagoons to the sea. These are broad on the windward eastern side of the atoll and narrow on the west. In total, the land area of Turneffe amounts to 10,831 ha. Added to this, some 400 ponds and lagoons within the islands occupy a further 252 ha. ' Department of Geography, The University of Edinburgh, Drummond Street, Edinburgh EH8 9XP, Scotland, UK Manuscript received 10 October 1998; revised 13 May 1999 Geologically, Turneffe lies on the second of five tilted fault-blocks that form sub- parallel, SSW trending submarine ridges. The faults are down-thrown towards the basin — dropping to the east (James and Ginsberg, 1979). The innermost ridge runs from Ambergris Cay in the north, down to the complex of cays off Belize City (St George’s Cay, Drowned Cays, etc.). The second ridge includes Chinchorro Bank (NE of Belize), Turneffe Atoll and the barrier reef islands off Dangriga (e.g. Tobacco Reef). The third ridge includes Glover’s Atoll and Lighthouse Reef. Thus, although on a map Turneffe appears very exposed to the east, it is actually shielded by three submarine ridges. The closest of these is the ridge forming Glover’s Atoll and Lighthouse Reef. The main surface circulation pattern shows water from the Atlantic flowing towards the west. Average surface flow velocities are thought to be in the range of 50-100 cm s' (James and Ginsburg, 1979). The mean wave direction can be summarised as a series of crests perpendicular to 074.5°N (Rutzler and Macintyre, 1982). This means that Lighthouse Reef will act to shelter the southern half of Turneffe, locally reducing wave velocities and favouring deposition. Many of the cays which form the Belizean Barrier Reef have been studied in detail (e.g. Woodroffe, 1995; Stoddart et al. 1982). Such works tend to focus on the more accessible cays of the inner reef, rather than the islands of the outer zone - Lighthouse Reef, Glover’s Atoll and Turneffe. The only previous examination of Turneffe as a whole was undertaken during the national resource assessment by Wright et al. (1959). Soils, vegetation and land use were recorded but the results were relatively cursory, given the agricultural emphasis of their work (Wright pers. comm. 1996). Furthermore, the resulting plant communities derived and shown on their 1:250,000 National Vegetation Map are somewhat misleading (Zisman, 1992). More detailed assessment of Turneffe’s terrestrial characteristics has been restricted to studies of the smaller cays of its eastern shore (Stoddart 1962, 1963). As a result, the soils and vegetation found on the larger islands have not been comprehensively surveyed. Over the period 1988-1993, Belize experienced rapid economic growth, a significant part of which was due to the increase in coastal tourism (World Bank 1996). As a result, Turneffe Atoll has come under close scrutiny from tourism and real estate developers seeking to establish resorts (Zisman 1998). In response, Belize's embryonic coastal zone management institution has sought new information on the atoll's natural resources. This need is reflected by the establishment of University College Belize’s Marine Research Centre at Calabash Cay in 1996/7 and the extensive underwater survey activities of Coral Cay Conservation Ltd. To assist in terrestrial resource evaluation, the authors were invited to carry out a rapid reconnaissance survey of the Turneffe Atoll. METHODOLOGY A combination of methods were used to investigate Turneffe's evolving land use, recent land cover and soil characteristics. Government archives yielded a range of secondary data that were used to reconstruct Turneffe's land use history. Satellite imagery, air photos and over-flights were deployed for land cover mapping and to locate a comprehensive range of survey sites. Field work was completed in April 1995, during the dry season, allowing examination of soils and vegetation. For a basemap, a preliminary outline of the cays was derived by tracing the islands from black and white 1:40,000 scale (1993) stereo aerial photographs and digitising this on to a GIS (Arc/INFO). The resulting map was geo-referenced using a network of 12 ground co-ordinates obtained using differential GPS. Fieldwork and scrutiny of 1990 air photos (also 1:40,000) allowed this outline to be refined. Land cover boundaries were mapped from these air photos onto acetate and then digitised for integration with the basemap. A false colour composite image of Landsat TM bands 3, 4, and 5 (scene 19/48, captured 4/1/1987) was produced to assist with air photo interpretation. This revealed differences in land cover poorly differentiated on the black and white air photos. The composite was also used to stratify the placement of sample points. This ensured that examples of all the vegetation communities were visited, and a broad geographical spread of points was obtained, reflecting the different geomorphological environments across the atoll. In total, 55 sets of observations were made at the 40 sample points selected (shown in Figure 2). At each point, field measurements were taken of interstitial soil-water pH, conductivity, sulphide and sulphate levels. The sample sites were also coded according to observed drainage and the energy level of the geomorphic setting. As well as providing information essential to the understanding of soil-plant relations, this information allowed soils to be classified according to the national framework for Belize, produced by Baillie et al. (1993). Structural characteristics of the different vegetation types were also recorded. The dominant plant species at each site were identified using the inventories of Fosberg et al. (1982) and comparison with specimens from the Forest Department and the Royal Botanic Garden, Edinburgh. Upon return to the UK, the provisional soil and vegetation community classifications were tested using dendrograms and clustering methods. Soil samples were grouped together using the Gower Similarity Index, Vegetation communities using Jaccard’s Index. Maps showing the resulting land cover classes were produced using Arc/INFO. LAND USE HISTORY Turneffe has a long history of settlement, dating back at least 1100 years. In order to understand present land cover, it is therefore necessary to identify the legacy of previous inhabitants on the atoll’s vegetation and physical features. This is the purpose of the following summary, drawn from secondary sources and field observations. Turneffe Atoll SOUTHERN LAGOON Grid squares are 10 km across i Projection: UTM (Zone 16) Spheroid: Clarke (1866) Units: Metres Figure 2 Location of the sampling sites As mentioned by Wright et al. (1959) and MacKie (1963), pottery fragments on several of Turneffe's cays indicate the presence of Maya Indians in the past (dating from the Late Classic to early Post-Classic). From their abundance, Wright er al. speculate that, as well as fishing camps, the Maya may well have established “more or less permanent trading posts with a permanent population engaged in a certain amount of subsistence farming” (p. 256). The likelihood of the prolonged existence of these trading and fishing settlements is made more probable by recent findings of such settlements on other cays in Belize (McKillop and Healey 1989). Lasting impacts on Turneffe's vegetation, soils and land cover are difficult to assess, but clearly, well-defined shell- heaps raised the level of many small areas and enabled colonisation by non-mangrove species. There is no evidence from remote sensing imagery or field work that the shoreline of any areas were artificially altered to create salt evaporation ponds, which have been found elsewhere in Belize. Colonial contact from the 1600s onwards led to the decimation and complete withdrawal of the coastal Maya from Turneffe. No evidence has been identified that precisely dates this event but the atoll was already charted, and therefore presumably visited, by the Spanish in 1625. The Mayan demise enabled regeneration of the atoll's vegetation as Colonial settlers were far less numerous and their encampments were generally transitory, particularly up to the late 18th century when fighting and attrition between British and Spanish forces ceased. As evidenced by the considerable detail of the West India Pilot of 1771, by this time, the British knew the area well and were frequent visitors (Stoddart 1962). Physical evidence from the era is lacking, but it is likely that camps were established on higher land for ship repairs and harvesting tortoise shell (Wright et al. 1959). Areas of fertile higher land were partially cleared for small farm plots, and from this time, fruit trees, goats and pigs were introduced to the islands (Jbid.). It is most likely that prime sites on the atoll were used, and that the initial alterations to these made by the Maya were compounded by early colonialists. The extent of settlements was still, however, relatively limited by the small number of people present. It is likely, therefore, that Turneffe's vegetation remained relatively intact. Towards the end of the nineteenth century, two new activities became established on Turmeffe, reflecting the general growth of the Colony’s economy. Between the late 1880s and early 1920s, the sponge industry flourished on the atoll, based on the “farming” of sponges in the southern inner lagoon (particularly Hippospongia gossypina, the “velvet sponge” - see Stevely and Sweat, 1994). Disease sealed the fate of this livelihood, however, wiping out the industry in 1919. Despite attempts to restock the lagoon in the 1920s and 1930s (Stoddart 1962; Smith 1941), sponge farming failed and left no significant impacts on the land cover °. 7 A 1938 map of the licensed sponge growing areas at Turneffe is held in the Public Record Office, Kew, London, UK. 6 In contrast to the sponge industry, the second activity established on Turneffe during the 19th century resulted in far more extensive impacts on the land cover than any other activity to date. This was the establishment of commercial coconut (Cocos nucifera) plantations. As noted by Stoddart (1962, 1963), coconuts were already present in 1720 on Lighthouse Reef, a second atoll 30 km further east. It is almost certain, given its proximity and the direction of currents, that they were present on Turneffe from this time as well. It was not until the latter part of the century, however, that commercial plantations (cocals) were widely established, producing whole nuts for export to the US. At its peak in 1914, national production averaged 6 millions nuts a year (Goodban 1952), almost entirely from coastal plantations. Large areas of beach thicket and coastal forest were therefore cleared, mainly on land leased from the government specifically for this purpose. From the export data available, the industry first became important around 1870, and thereafter plantations were progressively extended until 1920, by which time nearly all the suitable coastal and cay areas had been planted (Anon. 1920). During this time, Turneffe became one of the main producing areas, with extensive plantations on both its eastern and western sides. From the late 1930s onwards, despite increasing prices, national production went into decline, from an average for the decade of 4.4 million nuts to 2.7 million over the 1940s. Nonetheless, Goodban (1952) estimated that up to that year, “Turneffe, still had 1,000 acres (405 ha) under coconuts”. The tall varieties grown and the location of the plantations meant that they were extremely vulnerable to wind damage. Following the devastation caused by Hurricane Hattie in 1961 (Stoddart 1962), little interest was shown in rehabilitating the industry (Jenkin et al. 1976). Fieldwork carried out for this work shows that since then, although the plantations have generally fallen into a state of neglect, the groves of self- seeding coconuts persist, albeit now mixed with other plants. Land cover mapping does, however, reveal that cocals now occupy only a tenth of the area that they did in 1952, a reduction almost entirely attributable to the impact of Hurricane Hattie. The last of the longer-established land uses on Turneffe are fishing camps built by the crayfish (locally known as lobster) fishermen. These small sites are strategically placed for access to water, higher land and clear views over fishing grounds (to prevent theft of lobster pots) (Zisman 1998). Fishing of this nature began to attract significant numbers of seasonal settlers from the 1930s, when commercial exports to the US began. As Wright et al. (1959) describe, this attracted a few more families to seasonal residence on the cays. They also found that “many of these families have small cassava, yam and sweet potato plantations and the soils of the old Maya fishing sites are made use of almost without exception” (p. 256). This also led to the introduction of a small range of decorative plants. Notably however, the spread of the most problematic of these, Casuarina equisifolia (known locally as ‘Christmas tree’) has not been extensive. To date, it is limited to less than a dozen trees, mainly on the leeward side of the atoll. This is in contrast to other cays and other parts of the Caribbean coast, including Florida, where this pine has become a major ecological pest because of its low value for native wildlife and its tendency to out-compete certain native vegetation (Meadows, 1986). Turneffe’s fishermen are the only group responsible for on-going exploitation of native flora, albeit restricted to just four species. The palmetto Thrinax radiata and Acoelorraphe wrightii are harvested to make lobster traps sy Exploitation is extremely widespread and extraction was evident at almost all of the 40 sites visited. It results in thinning of the more accessible stands but as both species regrow rapidly, the vegetation composition is not being radically altered. The two other plant species being harvested are both mangroves. Rhizophora mangle and Laguncularia racemosa are cut for poles used to mark the location of lobster pots and for various minor construction purposes around fishing camps. Dead red mangrove is also used for cooking when butane gas is not available. This type of exploitation is less widespread, again restricted to the most accessible stands near to fishing camps. The most recent development on Turneffe is the rapid expansion of tourism. Up to the early 1990s, two small sports fishing establishments were the only resorts on the atoll. In operation since the 1970s, these facilities had relatively minor impact, occupying limited areas of mangrove (1.6 ha) and beach thicket (2.9 ha) most of which had been converted to cocal already. However, the rapid expansion of tourism towards the late 1980s put a premium on picturesque coast with white sandy beaches and access to good dive sites. As well as meeting these criteria, Turneffe also had the benefit of comprising mostly national land, cheap for developers to lease and disposable by Ministers as a form of political patronage. Following the change of government in 1989, tourism on Turneffe escalated, with the leasing of large areas to politically favoured developers (Zisman 1998). Specifically, 455 ha were awarded on Blackbird Cay in 1990, covering the largest area of high land on the windward side of the atoll. A further 56 ha was leased to a second developer on Calabash Cay. This rapid escalation not only caused concern amongst conservationists but also to the atoll’s fishermen, who saw tourism as a threat to their livelihoods from its potential damage habitats and water quality (/bid.). Given these increasing pressures, it has become important to (1) identify the habitats on the atoll, and (2) understand the processes that govern their composition and sensitivity to environmental change. The findings of this research are now presented in response to these requirements. The first sections detail the vegetation associations identified in the field. By combining these with Stoddart’s (1962) classes, a comprehensive classification of all the atoll’s plant communities has been assembled for the first time. : Acoelorraphe wrightii is now the accepted name for this species. Paurotis wrightii is an earlier synonym used by Stoddart (1962, 1963), and Wright et al. (1959). In the next section, the composition of Turneffe’s land cover is quantified (as of 1990), including the remaining extent of its natural plant communities. The third component of the research results is the soil analysis. Finally, consideration is given to the inter- relationships between vegetation, soil and land use, in particular, to the characteristics with immediate implications for proposed tourism development. IDENTIFYING THE VEGETATION TYPES FOUND ON TURNEFFE ATOLL As already stated, Stoddart (1962, 1963) produced an early vegetation classification for Turneffe. However, it was based on fieldwork almost entirely restricted to the atoll’s small eastern cays. Examination of the vegetation of the atoll’s larger islands was limited. The author therefore stresses the preliminary nature of the vegetation classes derived. For the present work, it was considered most appropriate therefore, to concentrate on surveying the larger islands, then to re-examine Stoddart’s classification and hence to produce a combined vegetation classification encompassing all of Turneffe’s vegetation types. A secondary consideration was that the classification should also be applicable to Belize’s other atolls, Lighthouse and Glover’s Reef. Field classification of the vegetation communities Six vegetation types have been identified during the reconnaissance survey: (i) mangrove, (ii) beach thicket, (111) broken palmetto thicket, (iv) broken palmetto- buttonwood thicket, (v) palmetto buttonwood scrub and (vi) cay forest. The term ‘broken’ refers to the local description of vegetation with an uneven, discontinuous canopy. It has been widely employed in previous descriptions of Belizean vegetation (e.g. Wright et al. 1959; Jenkin et al. 1976; King et al. 1992). Of the six vegetation types, mangroves are by far the most intensively studied. A national survey of Belize’s mangroves was carried out in 1991 (Furley and Ratter, 1992). Building on this work, a classification of Belizean mangrove sub-communities has been developed (Zisman 1998) based on differences in physiographic setting, species and vegetation structure. As the objective here is to establish only broad vegetation groupings, such a detailed breakdown is considered unnecessary. Thus, in this work the different mangrove communities are combined into a single category (although they are mapped in three separate height classes - tall, medium and dwarf). Mangroves The mangrove association varies widely, in composition and structure. At the level of highest structural development, it forms tall monospecific stands of black, white or red mangrove (up to approximately 18 m tall). Intermediate forms comprise mixed or pure red thickets, whilst the least structurally developed communities are the stunted black mangrove scrub and dwarf red mangrove. Overall differences in salinity, seasonal salinity fluctuations, inundation regime (depth, duration and frequency) and, nutrient inputs (particularly phosphorus and nitrogen availability) are responsible for this variation (Murray, 1995; McKee, 1993). The mangrove association is most common leeward of Turneffe’s main sand ridge, but mangroves also grow on the windward side, particularly where extensive reef flats reduce the wave energy nearshore. Fringing mangrove occur on the lee of the windward coral rubble cays, such as the Deadman Range. Mangroves also dominate the cays within the central lagoons. Mangroves are recorded growing on peat soils, sands and coral rubble. Beach thicket The beach thicket association is located on windward sand and coral rubble ridges. Such locations are relatively well drained, due to the underlying sandy or coral rubble soils. The structure of the beach thicket community is heavily influenced by exposure to the prevailing winds, in many cases this leads to a marked stunting. Broken palmetto thicket Broken palmetto thicket is found at a wide range of locations around Turneffe, both coastal and inland, both leeward and windward. It is a broad association, covering thickets of varying densities. The palm Acoelorraphe wrightii (known locally as “palmetto”) forms a significant component of the vegetation. Myrica cerifera, (“teabox”) is the other main species. The relative abundance of palmetto in this association is variable but insufficient to justify further division given the reconnaissance nature of this study. Suffice it to note that at four sites (numbers 25, 32, 51 and 52) a mangrove/broken palmetto thicket transition is found. This transition is a mix of mangrove species (R. mangle, A. germinans and the fern Acrostichum aureum) with terrestrial trees and shrubs from the broken palmetto thicket association. A second variant also exists which includes abundant Metopium brownei (‘black poisonwood’). It is thought that this indicates a transition to cay forest. Broken palmetto thicket occurs on drained peat and organic sand. Broken palmetto-buttonwood thicket The broken palmetto-buttonwood thicket association was isolated from the other thicket because of the predominance of Conocarpus erectus (‘buttonwood’), a mangrove associate (Tomlinson, 1986). In some areas buttonwood is found almost to the exclusion of palmetto. The association ranges in form from a dense thicket to a scrubby savanna, where “cutting grass” (Scleria spp.) occurs between clumps of C. erectus and shrubs such as Myrica cerifera. Broken palmetto-buttonwood thicket grows on drained peat, and evidence of burning suggests it may be a fire affected plagio-climax. Although likely to exist in smaller patches amidst the broken palmetto thicket, only one extensive area of this community has been identified and mapped, on the western side of the atoll. Palmetto-buttonwood scrub This community is also recorded from one major area only, also on the west of the atoll. Again, finer scale mapping and further investigation is likely to reveal areas of this 10 scrub associated with broken palmetto-buttonwood and broken palmetto thicket. Several sites presently supporting scrub show signs of recent burning (e.g. numbers 37, 38 and 53). It is thought that regular burning may be important in maintaining the characteristic open canopy of this class. It is this open canopy which differentiates it from broken palmetto-buttonwood thicket. Palmetto-buttonwood scrub favours sites with drained peat soil. Cay forest Cay forest is the climax association of the higher cays, where saline influence upon the groundwater is minimal or absent. It is restricted to sand ridge areas and organic sands. Structurally, it is the most developed vegetation association, forming fully developed closed canopy forests. Wright et al. (1959) record this type of vegetation on the higher more stable cays, suggesting that “at one time, mahogany and sapodilla forest were established on parts of Soldier, Calabash and Ropewalk Cays of the Turneffe archipelago” (p. 254). The present survey identifies only small isolated patches of this community, all of which have been altered to some degree, due to earlier farming. Of the four cay forest areas remaining in 1990, the largest is 13.7 ha on Blackbird Cay, part of the 455 ha parcel slated for tourist development. A comprehensive classification for Turneffe’s vegetation communities By combining the vegetation associations described above with those of Stoddart (1962), it is possible to provide a preliminary categorisation of all the vegetation types found on Turneffe. Some changes to his original nomenclature are necessary, however, to maintain consistency and for the sake of clarity. Stoddart’s sand-area thicket is renamed beach thicket and the broadleaf forest, cay forest. His other non-mangrove classes are retained. As already explained, his mangrove categories have been superseded and a single mangrove class is used here. Table 1 and Table 2 summarise the characteristics of the resulting vegetation associations, giving information on structure and dominant species. Turneffe’s vegetation has therefore to be sub-divided into ten basic associations. Some overlap is likely between beach thicket and marginal and interior thicket, but at the reconnaissance level the divisions are relatively satisfactory. The only other vegetation community identified were single species ‘meadows’ of Batis maritima but these were too few in number and too small (generally less than 0.01 ha) to warrant separate delineation. Testing this vegetation classification In order to test the vegetation classes developed in the field, nearest neighbour cluster analysis was applied to the sample site data upon return to the UK. This numerical 11 technique arranges the sample sites in a dendrogram so that sites with a similar suite of plant species plot close together. Following the algorithm protocol given by Dunn and Everitt (1982), similarity is assessed using Jaccard’s Index (Jaccard, 1912). This simple similarity coefficient is designed for use with binary (presence/absence) data. The resulting dendrogram is shown below in Figure 3. The grouping in the dendrogram accords well with the classes developed in the field. Five groups can be identified in the dendrogram, using the first branches at the left hand side of the figure, corresponding to a similarity index of zero (i.e. maximum difference). Examining the species present at each site in an individual branch reveals the “indicator species” used in the classification. From top to bottom, these groups are defined by the presence of: (i) Rhizophora mangle or Avicennia germinans, (ii) Metopium brownei and Thrinax radiata, (iii) Laguncularia racemosa, (iv) Cocos nucifera, and finally (v) Acoelorrhaphe wrightii and Scleria sp. The first group, defined by the presence of R. mangle and/or A. germinans, corresponds well with the mangrove vegetation class developed in the field. The second group, based around M. brownei and T. radiata comprises sites from cay forest and the poisonwood variant of broken palmetto thicket. The fact that these two field classes are combined in the Jaccard classification supports the idea that this poisonwood variant of the broken palmetto thicket is indeed a precursor to the development of cay forest vegetation. The third group, defined by the presence of L. racemosa contains all the sites classified in the field as beach thicket. In addition, it contains the mangrove site 36 (included in this group because, unusually, it contains a cover composed completely of white mangrove) and three other sites classified in the field as a possible transition. This transition class contains species representative of both the mangrove and beach thicket classes and it is suggested that this group marks the inland limit of sites at which mangroves can compete effectively with other species. The fourth group is defined by the presence of cocal (C. nucifera) and therefore consists of sites which have undergone the greatest human alteration. These range from small cultivated patches of fruit trees and palms to areas where the natural vegetation cover has been completely replaced by coconut plantations. The final group of sites on the dendrogram are linked by the presence of A. wrightii and S. bracteata. This category encompasses both the broken palmetto thicket and palmetto-buttonwood scrub classes devised in the field. This pairing supports the interpretation that palmetto-buttonwood scrub is a degraded form of the broken palmetto thicket. Both show a similar range of species, but the scrub has a lower density of trees and shrubs. The fact that the first branching separates mangroves from the other vegetation associations is likely to be due to the low range of species present at the sites covered by mangroves. The Jaccard index is known to be sensitive to species richness (van Tongeren, 1995). In general, however, the field and numerical classifications show a high degree of commonality. This suggests that they correspond with actual vegetation associations found at Turneffe. Furthermore, the placing of the less common field =" ?2 Table 1 Plant associations identified by the reconnaissance survey. Vegetation association Mangrove Beach thicket Cay forest pees te ts bed OP be Bp ie bt a eee Canopy: Bursera simaruba, Broken eee ee eos Canopy: Acoelorrhaphe wrightii, palmetto thicket Broken Ce et ee eee Canopy: Acoelorrhaphe wrightii, palmetto- buttonwood thicket Palmetto- buttonwood scrub Vegetation structure Height: wide ranging. Includes single species and mixed mangrove stands, from 0.6m high scrub to 18m high forest with variants between. Canopy: complete to sparse. Understorey: generally absent. Shrub layer: generally absent. Herb layer: present where canopy is open or moderately open, with scarce to abundant mangrove seedlings. Ground cover: leaf litter, algal mats and/or standing water. Height: approx. 2-7m depending on degree of exposure and stunting. Canopy: continuous to slightly broken. Understorey: generally absent. Shrub layer: generally absent. Herb layer: occasional woody herbs, plus grasses and xeromorphic species Ground cover: leaf litter or bare sand/coral rubble. Height: approx. 7-15m. Canopy: continuous. Understorey: present. Shrub layer: generally present. Herb layer: sparse or absent. Ground cover: leaf litter. Height: approx. 3-8m. Canopy: rarely continuous, generally slightly to moderately discontinuous. Understorey: generally absent. Shrub layer: present. Herb layer: sparse. Ground cover: leaf litter. Height: approx. 3-7m. Canopy: rarely continuous, generally slightly to very discontinuous. Understorey: generally absent. Shrub layer: present. Herb layer: continuous to sparse. Ground cover: bare ground, algal mats or leaf litter. Height: approx. 0.3-2m. Canopy: moderately to extremely discontinuous Understorey: absent Shrub layer: present Herb layer: continuous to sparse. Ground cover: leaf litter. Main species Canopy: Rhizophora mangle, Avicennia germinans, Laguncularia racemosa. Occasional Conocarpus erectus at transition to higher ground. Understory, Shrub & Herb layers: Occasional young of above, plus Batis_maritima. Canopy: Coccoloba uvifera, Cordia sebestana, Bursera simaruba, Cocos nucifera, Conocarpus erectus, Suriana maritima, Thrinax radiata and occasional mangroves. Understory & Shrub layer: Erithalis fruticosa, Pithecellobium keyense, Chrysobalanous icaco, Tournefortia gnaphalodes. Herb layer: Hymenocallis littoralis, Sesuvium portulacastrum, Wedelia, Stachytaphera, Andropogon, Cyperus and Eragrostis spp. Metopium brownei, Cordia sebestena, Thrinax radiata, Ponteria campechiana, Bumelia retusa, occasional Coccoloba uvifera. Understory: Ficus spp. plus young individuals of above. Shrub layer: Pithecellobium keyense. Herb layer: No data. Metopium brownei, Conocarpus erectus. Understory & Shrub layer: Myrica cerifera. Herb layer: Scleria bracteata. Conocarpus erectus, with occasional Metopium brownei. Understory & Shrub layer: Myrica cerifera. Herb layer: Scleria bracteata. Canopy: Acoelorrhaphe wrightii, Conocarpus erectus, Understory & Shrub layer: Myrica cerifera. Herb layer: Scleria bracteata. Stoddart’s equivalent classes Predominantly Rhizophora on mangrove cays. Mangrove transition zone on mangrove-sand cays. Rhizophora on mangrove-sand cays. Sand-area thicket on mangrove-sand cays. Broadleaf forest on sand cays. No equivalent as not covered by Stoddart. No equivalent as not covered by Stoddart. No equivalent as not covered by Stoddart. 13 Common names of these species are given in King et al. (1992) and Wright et al. (1959). Table 2 Additional plant associations described by Stoddart, 1962. Vegetation Typical Vegetation Main association setting structure species Strand Sand and Height: up to approx. 0.3m. Herb layer: Outer zone — Sesuvium mangrove- Canopy, Understorey & Shrub layer: not portalacastrum, Ipomea pes-caprae. sand applicable. Inner zone -Sporobolus virginicus, cays. Herb layer: sparse to continuous. Euphorobia spp., Canavalia rosea. Ground cover: bare sand. Other Sand Height: no data. Herb layer: grasses interior cays. Canopy, Understorey & Shrub layer: not areas - applicable. chiefly grass Herb layer: continuous to sparse. Ground cover: bare ground, algal mat or sand. Interior Sand Height: no data. Canopy: originally likely to be marsh and cays. Canopy, Understorey & Shrub layer: Avicennia germinans, Laguncularia swamp f¢. generally removed. racemosa, Rhizophora mangle, Herb layer: present. Conocarpus erectus. Ground cover: algal mat and/or standing Herb layer: Wedelia trilobata. water. Cocal Cocal on Height: approx. 7-15m. Canopy: Cocos nucifera. sand Canopy: continuous to moderately Herb layer: Stachytarpheta cays. discontinuous. jJamaicensis, Ambrosia hispida, Understorey: generally absent. Cakile lancelota, Wedelia trilobata. Shrub layer: generally absent. Herb layer: present especially under moderately discontinuous canopies. Ground cover: bare ground and/or litter of palm fronds. Source: Stoddart (1962) except for the vegetation structure data (this work). Common names of these species are given in King et al. (1992) and Wright et al. (1959). + Indicates that these vegetation associations are anthropomorphic. The interior marsh and swamp is created by clearance of basin mangrove sensu. Lugo and Snedaker, (1974), the cocal by clearance of cay forest, beach thicket and occasionally small areas of mangrove. derived classes (e.g. the poisonwood variant of the broken palmetto thicket) with other related groups (e.g. cay forest) corroborates the hypothesised pattern of vegetation succession. Thus, this analysis supports the validity of the field vegetation classification. Quantifying land cover on Turneffe Atoll The distribution of land cover across the atoll is shown in Figure 4 and quantified in Table 3. The map, produced using Arc/INFO GIS, is derived from interpretation of 1:40,000 black and white air photos (dating from February 1990 and March 1993). Reference was also made to 1987 Landsat TM imagery and ground survey at the points shown above. 14 eyI!4 | CNEWeY Ua01g JeVIIY) CNEWIEY UB01g WHI) CNEWIEY UExO1g mung poomuonng-onewyjey muss poomuoung-onewyjeg muds poomuonng-onewey UoIsuRIa qruds - 324214) ONEWIEY P14] ONEWEY U9404g O7DIY | ONEWIeY Ue}01g yey214] YoReg yeHP14] YIeeg @aosBueLy WHY | OHEW/EY USex}O1g JaHIIY | YoReg ysei04 Ae> yses04 Ae> (2214ym) aacuBueLy uon!sues) 294214] YDRag - eAcIBURLY uon{sues) 39x14) YDVeg - eACIBUELY yay214) Yeeg Jax214 | YDReG Jex214) Yoeag W@xI14) _YoReg uonisues) 9x2!14| YReeG - BACIBUELY (mq) 3@9P14), ONEWEY Uax0Ig ysasoy Ae> (Mg) 3®9P14 1 OCHEWeY Ua{041g (mq) 3@4714, ONEWIEY UexO1g yay214 YDREg (mq) 2@9PIYL ONAWIEY UEx04g (mq) 22xPIYL ONAWIEY UExO41g ysaio4 Ae> (mq) 2@4714, CNEWIEY Uax}OIg @aousueLy eaosZueLy @AosBueLy eaouBueP) @AojBueLy aAoiBuELy aaosBueLy @AouBueL) aaouzuepy aAosBueLy eaqusuepy aaou3ueLy aAouBueLy @AosBUeLY @aouBueLy BAosBueLY SSe/2 UONRIEZAA pjai4 | | | | So3 00’! $£0 0s°0 sv0 00'0 sot s eats ike! S AyseRjIWUIS JO Xapul Ss pserde[ Spas s s = ey S dg. @ s aa b 3) s a 64 s ges set c 2 a 2g p) Ss k 3 ce > so e) re a > God Eis 2) > 5 ga Oa sac © tod © cue & 41 | OHO 1 & ssc fed ep i D 2 & g 1 ®© Ca2'G 1 > “dos if a Lea 1 Q mF .% 9 Bure F 2 E88 Fi 9 Aas F > apo 2 jr oS) Povged 3 Q © 3M 4 E sa 33 4d coc Ve =| mM 4 A 4 & 287 © z 3 @ Th aS S — al e Ww Cy SE caik Ww = §Ss Ww © g-2 Ww S 630 & 26s W QQ) Sos o. 9 288 Ww CO eas Ww re) gS 3 S&S Ww “a oh 5 My xe@puy S,pseroe[ Buisn peunseaw Aq4eILUIS m 2s Be Ww siskjeue J93sN|> sNOquBiau saseayy © Be = te W Saha Ley aay (6) 1% Ww ,T fos =D weRs. groupings developed in the field. 15 The results show that 98.4 % of Turneffe is under natural (or near-natural) vegetation, with human activities responsible for the alteration of the remainder (176.5 ha). This is a considerably smaller area than earlier this century following the demise of coconut plantations, from the 405 ha in 1952 to 105 ha in 1990. The most significant finding in relation to contemporary changes in land use however, is the recent increase in tourism development. Mapping reveals relatively large areas of broken palmetto thicket and beach thicket degraded for tourist resorts. Over 48.1 ha was affected in this way, all between mid- 1989 and March 1990. Evidently, tourism is already having substantially increased impacts on Turneffe’s land cover. Mapping also reveals the limited extent of cay forest. Confined to 24.5 ha in 1990, it has been reduced by 46% from its original extent. This is of concern because it contains the highest biodiversity amongst any of the vegetation communities identified on the atoll. The information presented to date comprise Turneffe’s land use history, vegetation composition and land cover. The final component studied during fieldwork was its soils. Examination of this facet helps to understand land cover patterns, it also is useful for determining certain environmental implications of recent land use. The next section of this paper therefore reports on the investigations of soil properties at the 40 study sites. ANALYSIS OF SOIL CHARACTERISTICS Field examination of soil samples taken around Turneffe Atoll allows the development of a preliminary classification. This divides the soils into four groups based largely on differences in their organic content and drainage. This field classification accords well with the results of later testing using numerical agglomerative classification methods. These four field soil groups are then labelled according to the existing soil nomenclature employed in Belize, that of Baillie et al. (1993). Table 3 Nature and Extent of Land Cover on Turneffe Atoll, March 1990 Type of Natural Land Extent Anthropomorphic (modified) ~ Extent Cover Remaining ha Land Cover ha Mangrove 7419.3 Cleared mangrove 1.0 Degraded mangrove 0.4 Fishing camp in mangrove 10.1 Cocal in mangrove 3.6 Tourist area in mangrove 1.6 Marina/boat access in mangrove 0.6 Total Mangrove Altered 17.3 Beach Thicket 138.3 Cleared beach thicket 0.7 Degraded beach thicket 16.3 Cocal in beach thicket 77.3 Tourist area in beach thicket 4.5 Fishing camp in beach thicket 1.3 Total Beach Thicket Altered 100.6 Broken Palmetto Thicket 2303.1 Fishing camp in broken palmetto 1 187/ Thicket Total broken palmetto thicket 1.7 Altered Broken Palmetto Thicket 28.8 Degraded broken palmetto thicket 31.8 poison wood variant) poison wood variant) 41 Cocal in broken palmetto thicket Total broken palmetto thicket 35.9 poison wood variant) altered Broken Palmetto- 602.3 Buttonwood Thicket None altered 0.0 Palmetto-Buttonwood 314.4 Scrub None altered 0.0 Cay Forest 24.5 Degraded cay forest 0.7 Cocal in cay forest 20.3 Total cay forest altered 21.0 Figure 4 1990 Land Cover for Turneffe Atoll Key: (ee) Mangrove Cay Forest Beach Thicket os Broken Palmetto Thicket (I [[]) Broken Palmetto-Buttonwood Thicket ee Palmetto Buttonwood Scrub | Coconut Plantations S88 Other modified land cover Degraded and cleared vegetation 1940 000 [= ]| NORTH SHEET N A Scale 1:130000 000 Ly 1930000 == [== Broken Palmetto Thicket (Poisonwood variant) aN a neve ocennecenenconsconcen/ecrcastescctoaoncaoensceieccaceiectaecovaronssosaeegnsetsessseettiset i 000 cr Projection UTM (Zone 16) pheroid Clarke (1866) 18 Figure 4 1990 Land Cover for Turneffe Atoll 1910 000 Scale a/9001G00 1:130000 SOUTH SHEET Projection UTM (Zone 16) pheroid Clarke (1866) 19 Classification of soils in the field A four-group soil classification was developed in the field, namely: (i) waterlogged saline peat, (ii) drained peat, (ili) organic sand, and (iv) coral sand. Waterlogged saline peats are deep, permanently wet organic soils. They are found across the full range of low to high energy environments around Turneffe, on both the leeward and windward sides of the islands. Drained peats are seasonally dry, highly organic soils. They are found in low to medium energy environments on both the leeward and windward sides of the islands and along the shores of lagoons. Organic sands are well drained, moderately organic soils. They are found in all geomorphic settings around the atoll, but are most common on moderate and high energy, windward coastal settings. Coral sands are well drained carbonate-derived mineral soils, with clasts which vary in size from sand to cobble fractions. They are found almost exclusively in high energy, windward environments. The soils of Turneffe can effectively be split into two groups — the mineral sands and the organics. The mineral sands are derived from in situ weathering of terrestrial material such as “beach rock” together with wave and wind deposition of sand and coral rubble from offshore sources. The organic rich waterlogged saline peat has accumulated in areas supporting mangrove forest. In common with other coastal areas in Belize, it is composed largely of decomposing vegetative material from Rhizophora mangle plants. The origin of the drained peat is more problematic. To the naked eye it appears very similar to the mangrove peat, albeit slightly finer. Because of its present elevated position, decomposition of the drained peat is dominated by aerobic rather than anaerobic pathways and it too supports soil macrofauna. These will both act to reduce the average particle size. It is thought that the drained peat is indeed derived from mangrove areas. This requires a mechanism of uplift to explain its present elevation. The most likely candidate is storm wave and/or hurricane transportation. The banks of the creeks and bogues around Turneffe are composed of peat. Many of these are aligned east-west, i.e. along the track of previous hurricanes. The peat is highly fibrous and it is thought likely that blocks torn off would retain their shape rather than fall apart. Light and able to float, these peat “rafts” could be transported considerable distances. Further palaeo-ecologic investigation (such as pollen and diatom identification) are, however, needed to fully resolve the origin of this soil. Selected soil-water properties characteristic of the four identified soil groups are given in Appendix 1, and summarised in Table 4. 20 Table 4 Summary of the field soil class properties. Field Soil Vegetation Geomorphic Drainage Salinity pH classification association setting Waterlogged saline Mangrove All very poor high neutral to peat acidic Drained peat Broken palmetto thicket, low to slightly moderate neutral broken palmetto-buttonwood moderate impeded thicket Organic sand Cay forest, cocal, beach thicket, All good low slightly broken palmetto thicket alkaline Mangrove All moderate high slightly to poor alkaline Coral sand Beach thicket high energy good moderate slightly alkaline Cocal high energy good very low slightly alkaline Testing the soil classification The use of numerical methods to test classifications developed in the field is becoming increasingly common in soil science, mirroring their use in the wider field of ecology. These use algorithms to mathematically “cluster” sample sites according to a numerical measure of similarity. This creates discrete groups of samples. One such measure considered applicable to studies of soils (Webster and Oliver, 1990) is the Gower Similarity Index (Gower, 1971). This measure was applied to the sample data using the program MVSP Plus (Kovach, 1993) and the resulting classification expressed pictorially as a dendrogram (Figure 5). Gower’s Index values were calculated for each sample site using both the measured soil parameters and scorings representing the drainage and the energy level of the site’s geomorphic setting. The sites were arranged into groups using an unweighted pair clustering algorithm. The dendrogram shows that the sites can be split into a number of groups, depending on the threshold value of the index used. At the extreme left of the figure, (index = 0.51) the sites can be allocated to one of two distinct soil groups; at the far right, (index = 1.00) 22 groups can be identified. The four-group classification developed in the field lies somewhere between these two extremes. Looking at the sample sites, it seems that the major factor separating the two largest (left-hand) groups is soil drainage. The upper 39 sites (later assigned to classes WP, Mix and OS) represent sites where for a large part of the year the water table is regularly very close to the soil surface. The lowest 10 samples are all typically well drained. As the number of classes increases, other differences are brought into play. 21 PUES [EIO=) Pires [et05) ajduues sare, Pures 21ue3IO ajdures sare pues (es0D Pures 21Ue319, pues 21ue81Q Teag peurg pues 21Ue81Q) pues 21U2810 pues 21Ue81Q) pues 21UeE, purs 2|Uue8uG pues 21ue3I pues 21Ue81Q, Tea paBBoparenyy purs 21ue21Q) Teed passoaieyy pues 21ue81Q, ajdures sareyy purs 21Ue319 purs 21ue2uO Teag passopinen pues 21UeB1O Teeg pasBoparep Teag pasZojsareyy qeag pessoa pues 21U2810, Rag PeureIg) pues 21Ue84Q) kag pasBojiaey pues 21uedO Teag passopaiey, Teag passojiaien Teag passojsaieyy Ieag paBsojsaien, ajdures sare,a, Akad pasBojiaren Tea passopiarey Ikag PasBojiaey SSE/P |!0S pjaly aus X@PU] S$ JBMO BuIsn painsea AWIe|ILUIS sisAjeue J03sn|> dnous sed payysiIamMuA X@PU] S$, JaMO5 Figure 5 Classification of sampling sites by soil type This figure shows the sample sites (numbered on the right) and their associated field soil descriptions (the nearby text) arranged in a dendrogram. The clustering of sites into groups has been achieved using a numerical classification routine. Five classes (labelled as WP, Mix, OS, DP and CS) can be distinguished. These classes are shown separated by dotted lines. 22 Moving to the right of the dendrogram, the number of possible classes increases. Figure 5 shows the dendrogram split into five classes (labelled WP, Mix, OS, DP and CS) at an index value of approximately 0.60. Four of these groups (WP, OS, DP and CS) appear well differentiated (for example, seven out of the nine sites in the WP class have waterlogged peat soils) and correspond closely to the classes identified in the field. The fifth group (shown as Mix on Figure 5) contains sites with soils of either waterlogged saline peat or organic sand. This fifth numerical class appears to be an amalgam of the others, rather than a further, equally unique group. It is probably an outcome of the limited suite of soil measurements taken. If this was broadened (e.g. to include measurements of the macro- nutrient elements) this artificial grouping would be expected to disappear. Soil properties from sites in this fifth class are not thought to differ significantly either chemically, or in the processes forming and maintaining them. Therefore, the addition of a fifth class to the field groupings was rejected. In any classificatory scheme there is likely to be some degree of overlap between the classes. The strong overall correlation between the field and numerical grouping of the sites supports the validity of the field classification. The four-group classification is therefore retained. Existing classificatory schemes for the soils of Belize Many schemes exist for classifying soils. The two international systems are ‘Soil Taxonomy’ (USDA, 1975; SMSS, 1990) and the FAO/UNESCO ‘Legend of the Soil Map of the World’ (FAO/UNESCO, 1974, 1988). The authors agree with Baillie et al. (1993) who have shown both these systems to have considerable drawbacks to applications in the humid tropics and Belize in particular. They criticise the low weighting given to parent material and the morphology of the entire profile in the ‘Soil Taxonomy’ system. Furthermore and significant to local use, Belizean farmers, forest workers and planners already use a system of nomenclature based on a 1959 survey by Wright et al. Many later local soil maps also employ this scheme. Any attempt to introduce the FAO or USDA schemes is therefore likely to increase confusion rather than clarity. The local terms are largely retained in the Land Resource Assessment of Belize produced by King et al. (1986, 1989, 1992) which provides a comprehensive mainland soil survey, mapping units at a scale of 1: 100,000. It is an example of the emergent Belizean three tier system of soil classification. This draws on the earlier system of Wright et al. (1959) and has since been consolidated and revised by Baillie et al. (1993). This, the local system, was chosen for use in this work. In the Belizean system, the soils are divided into suites, sub-suites and series. Soil suites are defined predominantly by their parent material, although some developed on similar lithology may be further differentiated by climatic differences. Sub-suites are defined 23 using major morphological and chemical features, with an emphasis upon agronomic utility. Because of the reconnaissance nature of the work of King et al., the soil classification was not extended to the series level. The study had an understandable mainland focus, meaning that few of the defined suites occur in the Turneffe area. However, one of the soil suites is relevant to the present work - the Turneffe Suite, first defined by Wright et al. (1959). A brief description of this soil suite, taken from King et al. (1992, p 72) is given below: Turneffe Suite The Turneffe Suite consists of well- and moderately drained. soils formed on Pleistocene to recent coastal deposits, sometimes underlain by shallow coral, and which can be predominantly siliceous or calcareous. Five sub-suites are distinguishable: Shipstern, Ambergris, Hopkins, Matamore and Barranco *. Full soil sub-suite descriptions are given in Baillie et al. (1993), but their key characteristics are summarised in Table 5 below. Table 5 Characteristics of soils from the Turneffe Suite. Parent Pedogenic Soil Agricultural materials environment characteristics potential Shipstern Rawshallowsandy and Coastal flats, near § Pale-coloured, some Limited. muddy sediments sea level, often soils are saline Shallow and droughty usually calcareous over _— covered by recently emergent coral. mangrove savanna. Ambergris Rawpale coarse soilsin Close to beaches. Raw, coarse, pale Moderate: cashew, deep calcareous sands sand, derived from pineapple and of modern or recent coral and/or shells. coconuts. beaches. Tourist potential Hopkins Raw pale coarse soils in Close to beaches. Raw, coarse, pale Limited: possibly deep quartzose sands siliceous sands. cashew. Droughty and of modern or recent infertile. beaches. Tourist potential. Matamore Young and slightly Inland, stranded Weathering leads to Moderate, for developed pink or yellow fossil beach yellow and yellowish- chemically sands on relict deposits. red colouring, weakly undemanding, deep, subrecent coastal developed fine texture, _ free-draining root-zone deposits. but still predominantly crops, e.g. cashew, coarse. coconuts and cassava. Barranco Moderately leached red No data* No data No data and yellow loams and clays on old coastal depoits. Parent material description are from Baillie et al. (1993, p43). Other descriptions are adapted from King ef al. (1992 pp187-213). * The Barranco sub-suite is not listed in King’s 1992 report. These sub-suites can be distinguished by differences in parent material and pedogenic * The Barranco sub-suite was added to the soil schema by Baillie et al. in 1993. 24 environment. Parent material differences divide the sub-suites in two. Soils with a high carbonate content (the Shipstern and Ambergris suites) have a marine source of minerals. Soils rich in silica such as the Hopkins, Matamore and Barranco sub-suites, are derived from terrestrial sediments. Therefore, the sub-suites found at Turneffe will be the carbonate dominated ones: Shipstern or Ambergris. Choosing between these sub- suites depends upon the pedogenic environment, primarily the depth of soil and the local depositional regime. Placing the soil groups identified in the field into the Belizean classification Relating the field identifications to the Belizean system, two of the field soil groups correspond closely with soils of the Turneffe Suite. The coral sand matches the Ambergris sub-suite and the organic sand is very similar to the Shipstern sub-suite, although the profiles found at Turneffe are generally deeper than those described by Baillie et al. (1993) and King et al. (1992). The high organic content of the waterlogged saline peat soil does not however, accord well with reported organic values across the Turneffe Suite. Instead, it is more typical of the Tintal suite (described by King et al. (1992, p 208). Tintal Suite Tintal Suite soils are poorly drained for all or a considerable part of the year. Soil formation is dominated by the gleying processes associated with wet and reducing conditions.... They occur in low-lying positions. This classification is commonly applied to terrestrial environments in Belize: wet alluvium and hillslope-material derived soils. Yet the presence of peat, the high salinity and the continual waterlogged state of the waterlogged saline peat soils accords with a sub-suite of the Tintal suite - the Ycacos (described in Table 6). Table 6 Characteristics of soils from the Ycacos sub-suite Tintal Suite) Parent Pedogenic Soil Agricultural materials environment characteristics potential Ycacos Permanently Coastal and Pedogenesis has hardly Negligible. waterlogged portion of pericoastal begun. Dominant chemical Poor drainage, mineral and organic swamps and small characteristic is its salinity: flood hazard and material. Very young — patches inland fed moderate to extremely high; __ salinity. parent material is still by saline or Calcium and magnesium are often accumulating. brackish springs. the dominant exchangeable cations. The only soil which does not seem to fit well in any of the established classes is the drained peat. It is suggested that this could be considered as a new series of the Ycacos sub-suite, as a result of its subsequent drainage and salinity. These findings are summarised in Table 7 below. 25 Table 7 Summary of soil classes found at Turneffe Atoll ___ Field Classification ____—~«Equivalent in the Belizean Classification system ——sSsssts=~—SsS Coral sand Turneffe suite Ambergris sub-suite Organic sand Turneffe suite Shipstern sub-suite Waterlogged saline peat Tintal suite Ycacos sub-suite Drained peat Tintal suite Possibly a new series in the Ycacos sub-suite The identification of key environmental gradients Ordination analysis (to be reported elsewhere) reveals two environmental gradients acting upon the vegetation. The first axis aligns sites along a well-drained-waterlogged soils gradient. The second differentiates between sheltered, low-energy, acid soils and exposed, high energy alkaline soil conditions. These gradients can be used to interpret the position of the clusters of sample sites which indicate vegetation communities. DISCUSSION OF PLANT-SOIL RELATIONSHIPS The controlling factor for mangroves is evidently the capacity to endure saline or brackish waterlogging, made possible by their characteristic morphological adaptations (pneumatophores and aerial roots). Sample sites covered by mangroves plot at positions near the centre of the second, acidity-alkalinity gradient. This is not to say that mangroves cannot tolerate either of the extreme conditions of acidity/alkalinity or high/low energy environments, for indeed mangroves are the plants with the widest physiological tolerance range of any found at Turneffe. Rather, it is precisely this wide tolerance to such conditions which means that this gradient does not act to constrain their colonisation activity. Broken palmetto thicket and broken palmetto-buttonwood thicket are found in sheltered, low-energy areas with well-drained soils. Lack of morphological adaptations means that they cannot tolerate soils likely to experience extended periods of flooding or survive saline or brackish influence. Well-drained conditions are also favoured by the beach thicket and cay forest communities, but their preference for soils derived from limestone places them at the alkaline end of the acid-alkali gradient. Beach thicket species are also adapted to extreme drought and low soil nutrient levels. This is not the case with cay forest species. The presence of sites covered by cocal is not the result of physiological restraints, for they have a wide range of tolerances (Langdon, 1991). Rather it is the 26 result of human activity, clearing beach thicket and cay forest areas to create coconut plantations. CONCLUSIONS Vegetation distribution The vegetation of Turneffe Atoll has been shown to be far more varied than the original “mostly mangrove” inferred from earlier maps such as that of Stoddart (1962) and Wright et al. (1959). Mangroves dominate low-lying areas, forming overwash islands or fringes along creeks, around lagoons and the coastline. In freely-drained windward areas cays are covered in beach thicket. In well-drained soils with a greater organic component, broken palmetto thicket, broken-palmetto buttonwood thicket or even cay forest occurs. In disturbed areas there are patches of cocal and palmetto buttonwood scrub. With the exception of the cay forest (which dominates the higher sand ridges), the vegetation of Turneffe is characterised by a low structural complexity and the absence of “climax” species. This is typical of coastal forests across the Caribbean (Roth, 1992). This low complexity can be attributed to the lack of nutrients and impact of hurricanes. Signs of these (both sites which have suffered considerable erosion and others which have experienced subsequent re-deposition) can be found across Turneffe. Hurricanes tend to encourage plants with a cyclical pattern of development. Mangrove species are characteristic of these conditions. Features such as the large number of propagules produced, sharp zonation and even-aged stands reflect their adaptation to a rapid cycle of growth and mortality (Jimenéz et al., 1985). Yet as noted above, mangroves are not the only species present. The considerable vegetation diversity which exists across Turneffe Atoll today (only a few decades after the last major hurricane) demonstrates that through seed transportation mechanisms (air, water and bird vectors) the other plant species are also able to re-establish themselves. Hydrology With the absence of terrestrial freshwater drainage features, rainfall is the only freshwater input on the cays. Their surface is typically composed of loose, open sediments which favour high infiltration rates. Surface runoff is likely to be very rare, if it occurs at all it will be limited to extremely large precipitation events. Plants will therefore be dependent upon a mixture of direct interception of rainfall during the wet season and groundwater sources in the dry season to obtain the water necessary for their survival. Urish (1991) has shown that cay size has a considerable effect upon both the volume and seasonal availability of fresh water lenses. He found that on small cays (such as Carrie Bow Cay — 0.4 ha) the fresh water lens effectively disappeared during 27 the dry season, to be replaced with a shallow brackish water zone. He notes that the small tidal range typical of Belize reduces the tidal effects. Thus, at Turneffe, freshwater lenses are likely to exist even on the smaller cays, which in areas with a greater tidal amplitude, would quickly be dispersed. This freshwater availability will act to favour species diversity, supporting non-halophytic species. In particular, given their deeper roots and sensitivity to brackish conditions, freshwater lenses are extremely important for the development and survival of Turneffe’s cay forest. Soils The soil chemistry and plant-soil relationships at Turneffe are broadly similar to those seen on other cays and calcareous-dominated mainland sites in Belize (e.g. Furley et al., 1993; Murray, 1995). The key differences relate to variations in soil salinity, redox potential and organic content. All soils are likely to show limited amounts of available nitrogen and phosphorus. Development In the four hundred years after the Maya were dislocated from Turneffe, human impact on its land cover was minimal. The establishment of commercial coconut plantations in the 1890s changed this, producing the most extensive habitat alteration ever to take place. At least 450 ha were cleared, mostly beach thicket but also significant areas of cay forest. However, with the declining viability of coconut exports and the devastation caused by Hurricane Hattie in 1961, many coconut groves began returning to semi- natural vegetation. Nonetheless, 105 ha still remain, by far the most extensive artificial land cover on Turneffe. Between the 1960s and late 1980s, land use on Turneffe remained relatively stable, and as a consequence, further human impact was minimal. However, evidence of land clearing and land speculation signals the spread of tourism since then. This development is opening a new chapter in the evolving exploitation of the atoll and already land clearance has rapidly escalated. This has caused considerable concern amongst the resident fishing community, who see risks to their livelihoods resulting from marine pollution and mangrove clearance. Conservationists are also concerned, since the atoll is a stronghold for the endangered Salt Water Crocodile (Crocodylus acutus), and may provide suitable sea turtle nesting habitat if disturbance is prevented. Furthermore, the recent discovery of the endemic gecko (Phyllodactylus insularis) on Belize’s other atolls, makes it likely that it is also present on Turneffe. To date, these are the only localities from which it is recorded anywhere in the world. The particular concern with tourism development is that it is concentrated along the coast and highest land (especially where freshwater wells can be sunk). Clearance for 28 the resorts themselves, for ancillary facilities (including a proposed airstrip) and for holiday homes makes it inevitable that mangroves and cay forest will be affected. Whilst the impact on the first is to some extent, mitigated by the prevalence of this habitat, it is not the case with cay forest. Now identified as the most floristically diverse habitat (and the environment likely to prove richest in archaeological interest), the remainder of this habitat needs protection and sensitive management. In addition, areas of beach thicket are being cleared. Far less damage to this habitat would result if existing beach front coconut plantations were used for tourism. In relation to the soil characteristics identified, the risk of accelerated erosion following the clearance of fringing mangroves is already well known. It is also evidently the case however, that the atoll’s dried organic soils are at risk from deflation and erosion if cleared of all vegetation. Belize’s embryonic coastal management bodies are trying to balance different stakeholder interests by promoting land use zoning for Turneffe (see McGill 1996). To date, they lack the legal framework or political backing to enforce the resulting plan, a fact explained by the links between resort developers and the ruling party (Zisman 1998). With a new government in place (September 1998), the sustainability of development on Turneffe is in the balance. ACKNOWLEDGEMENTS The first phase of this research was funded by the Darwin Initiative of the UK’s Department of Environment and was carried out on contract to Coral Cay Conservation Ltd. (CCC). Their staff in London and Belize City offices provided logistical support and materials. Later work was supported by the Carnegie Trust for the Universities of Scotland. Whilst on Calabash, CCC volunteers helped in many small, but much appreciated ways. Thanks are especially due to Kevin Coye for guiding us around the islands and contributing to the data collection, Expedition Leader Jim Thomas and Science Officer Paul Scott. University College Belize provided office space and computing facilities at their campus in Belize City and offshore accommodation at the Marine Research Centre. The Belize Center for Environmental Studies and the Belize Audubon Society provided valuable access to their library resources. The Forest Department allowed us to use their herbarium. The Lands and Surveys Department granted access to their aerial photograph collection. David Stoddart and Ian Macintyre made helpful comments on an earlier draft of this manuscript. Sincere thanks are extended to the individuals and institutions concerned. 29 REFERENCES Anon. (1920) Colonial annual report. Colonial Office, London, UK. Baillie, 1.C., Wright, A.C.S., Holder, M.A. and Fitzpatrick, E.A. (1993) Revised classification of the soils of Belize. NRI Bulletin 59. Natural Resources Institute Chatham, UK. Dunn, G. and Everitt, B.S. (1982) An introduction to mathematical taxonomy. Cambridge University Press, Cambridge, UK. Food and Agricultural Organisation of the United States/United Nations Educational, Scientific and Cultural Organisation - FAO/UNESCO (1974) Soil Map of the World. Vol. 1. Legend. Paris. UNESCO. Food and Agricultural Organisation of the United States/United Nations Educational, Scientific and Cultural Organisation - FAO/UNESCO (1988) Soil Map of the World. Revised Legend. Rome. FAO. Fosberg, F.D., Stoddart, D., Sachet, M.-H. and Spellman, D. (1982) Plants of the Belize cays. Atoll Res. Bull., No. 258. Furley, P.A, Minty, C.D., Murray, M.R. and Ross, S.M. (1993) Soil and sediment characteristics within coastal and inland mangrove communities. p. 34-78 In Furley, P.A. and Munro, D.J. eds. The wetlands of Belize: ecology, environment and utilisation. Department of Geography, The University of Edinburgh, Edinburgh, UK. ISBN 0 900475 22 6. Furley, P.A. and Ratter, J.A. eds. (1992) Mangrove distribution, vulnerability and management in Central America. ODA/OFI Forestry Research Programme Contract No R4736. Department of Geography, The University of Edinburgh, Edinburgh, UK. Goodban, J. W. D. (1952) A report on the coconut industry of British Honduras. Director of Agriculture, Belize. Gower, J.C. (1971) A general coefficient of similarity and some of its properties. Biometrics, v. 30, p. 643-654. Jaccard, P. (1912) The distribution of the flora of the alpine zone. New Phytologist v11 pp 37-S0. James, N.P. and Ginsberg, R.N. (1979) The seaward margin of Belize barrier and atoll reefs. The International Association of Sedimentologists, Special Publication No. 3. 191p. 30 Jenkin, R.N., Rose Innes, R., Dunsmore, J.R., Walker, S.H., Birchall, C.J. and Briggs, J.S. (1976) The agricultural development potential of the Belize Valley Land Resources Division, Tolworth, UK. F Jimenéz, J.A., Lugo, A.E. and Cintron, G. (1985) Tree mortality in mangrove forests. Biotropica, v. 17 p.177-185. King, R.B., Baillie, LC., Abell, T.M.B., Dunsmore, J.R., Gray, D.A., Pratt, J.H., Versey, H.R., Wright, A.C.S. and Zisman, S.A. (1992) Land resource assessment of northern Belize. Natural Resources Institute Bulletin 43, 2 volumes. Kovach, W.I. (1993) MVSP Plus - A multivariate statistical package for IBM-PCs, ver 2.1. Kovach Computing Services, Pentraeth, Wales, UK. Langdon, J.R. (1991) The Booker Tropical Soil Manual. Longman Scientific and Technical. New York. Land Information Centre/Coral Cay Conservation - LIC/CCC (1995) Geometrically corrected map of Turneffe Atoll. Land Information Centre, Lands and Surveys Department, Ministry of Natural Resources, Belmopan, Belize. MacKie, E. (1963) Some Maya pottery from Grand Bogue Point, Turneffe Islands, British Honduras. Cited In D. Stoddart Effects of Hurricane Hattie on the British Honduras Reefs and Cays, October 30-31, 1961. Atoll Res. Bull., No. 95, p. 131-135. McGill, J. (1996) Draft Comprehensive development plan for Turneffe Atoll. UNDP/GEF Coastal Zone Management Project and the Fisheries Department. Belize City, Belize. Mckee, K.L. (1993) Soil physiochemical patterns and mangrove species distribution - reciprocal effects? Journal of Ecology, v. 81, p. 477-487. McKillop H. and P. Healey (1989) Coastal Maya trade, Trent University Occasional Papers in Anthropology 8, Peterborough, Canada. Meadows, M. (1986) Casuarinas ‘pine trees’ in Belize. Belize Audubon Society Newsletter, v. 182., p. 3. Belize City, Belize. Murray, M.R. (1995) The environmental effects of mangrove clearance in Belize, Central America Unpublished Ph.D. thesis. Department of Geography, The University of Edinburgh. Roth, L.C. (1992) Hurricanes and mangrove regeneration: effects of Hurricane Joan, October 1988, on the vegetation of Isla del Venado, Bluefields, Nicaragua. Biotropica, 31 v. 243, p. 375-384. Rutzler, K. and Macintyre, I.G. (1982) eds. The Atlantic barrier reef ecosystem at Carrie Bow Cay, Belize: I Structure and communities. Smithsonian Contributions to the Marine Sciences v. 12, 539p. Washington D.C., USA. Soil Management Support Services - SMSS (1990) Keys to soil taxonomy. Blacksburg, Virginia. Virginia Polytechnic Institute and State University. Stevely, J.M. and Sweat, D.E. (1994) A preliminary examination of the commercial sponge resources of Belize with reference to the location of the Turneffe Islands sponge farm. Atoll Res. Bull. No. 424, 21pp. Stoddart, D.R. (1962) Three Caribbean atolls: Turneffe Islands, Lighthouse reef, and Glover’s reef, British Honduras. Atoll Res. Bull., No. 87, 151p. Stoddart, D.R. (1963) Effects of Hurricane Hattie on the British Honduras Reefs and Cays, October 30-31, 1961. Atoll Res. Bull., No. 95, 142pp. Stoddart, D.R., Fosberg, F.R. and Spellman, D.L. (1982) Cays of the Belize Barrier Reef and Lagoon. Atoll Res. Bull., No. 256, 76pp. Tomlinson, P.B. (1986) The Botany of Mangroves. Cambridge University Press, New York, USA. United States Department of Agriculture - USDA (1975) Soil Taxonomy. Soil Conservation Service, Washington D.C. USA. Urish, D.W. (1991) Hydrogeology of Caribbean coral reef islands. pp136-148 in Cambers, G. (editor) Coastlines of the Caribbean. American Society of Civil Engineers, New York, USA. 187 pp. van Tongeren, O.F.R. (1995) Cluster analysis. p. 174-212 In Jongman, R.H.G., ter Braak, C.J.F. and van Tongeren, O.F.R. Data analysis in community and landscape ecology. Second edition. Cambridge University Press, New York, USA. Webster, R. and Oliver, M. (1990) Statistical method in soil and land resource survey. Oxford University Press, Oxford, UK. Woodroffe, C.D. (1995) Mangrove vegetation of Tobacco Range and nearby mangrove ranges, Central Belize Barrier Reef. Atoll Res. Bull. No. 427, 35pp. World Bank (1996) Belize Environmental Report. Report No. 15543-BEL. Caribbean Division, Country Department II, Latin America and Caribbean Region. Washington 32 DG, Wright, A.C.S., Romney, D.H., Arbuckle, R.H. and Vial, V.E. (1959) Land in British Honduras. Report of the British Honduras Land Use Survey Team. Colonial Research Publications No. 24, Colonial Office, HMSO, London, UK. Zisman, S.A. (1992) Mangroves in Belize: their characteristics, use and conservation. Forest Planning and Management Project. Forest Department. Belmopan, Belize. Zisman, S.A. (1998) Sustainability or status quo: An assessment of elite influence in the political ecology of Belizean mangroves. Unpublished Ph.D. thesis. Department of Geography, The University of Edinburgh. APPENDIX I: SOIL DATA Table 8 Selected characteristics of the soils of Turneffe Atoll pH Mean Mode Range of _ values Waterlogged saline peat 7.24 6.87 5.56 to 8.58 Drained peat 7.20 - 6.18 to 7.96 Organic sand 7.56 - 6.80 to 8.06 Coral sand 7.68 - 7.56 to 7.85 Conductivity (mS cm” at 25°C Mean Mode Range of _ values Waterlogged saline peat 95.37 - 54.60 to 113.98 Drained peat 20.17 - 11.36 to 36.38 Organic sand 56.19 - 1.22 to 154.27 Coral sand 25.43 - 2.68 to 70.74 Sulphate (mg!) Mean Mode Range of _ values Waterlogged saline peat 1480 1600 400 to 1600 Drained peat 1275 1600 300. to 1600 Organic sand 931 1600 200 to 1600 Coral sand 733 300 300 to 1600 “Sulphide (mg) Mean Mode Range of values _ Waterlogged saline peat 3.5 0 0 to 10 Drained peat 1.2 0 0 to 5 Organic sand 1.2 0 0) to 10 Coral sand ¢) 0 0 to ¢) Water table height (cm Mean Mode Range of _ values Waterlogged saline peat -4.0 0.0 -240 to 20.0 Drained peat -43.8 -30.0 -98.0 to -30.0 Organic sand -58.9 : -150.0 to -17.0 Coral sand -81.6 : -150.0 to -28.0 Heights are expressed relative to the local ground surface. Negative values indicate that the water table was below the ground surface, positive values that standing water was present at the time of sampling. Absence of mode values indicates that the frequency histogram was not unimodal. ATOLL RESEARCH BULLETIN NO. 465 A MICROBIALITE/ALGAL RIDGE FRINGING REEF COMPLEX, HIGHBORNE CAY, BAHAMAS BY R. PAMELA REID, IAN G. MACINTYRE AND ROBERT S. STENECK ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. AUGUST 1999 BAHAMAS ELEUTHERA ————> {-“<— NEW ae PROVIDENCE 24°44' Fringing reef a 24°43' 24°43' Stromatolites in back reef lagoon HIGHBORNE CAY 76°50! FIGURE 1. Map showing the location of Highborne Cay, Bahamas; stromatolites form in the southernmost kilometer of a microbialite/algal ridge fringing reef complex along the east shore. A MICROBIALITE/ALGAL RIDGE FRINGING REEF COMPLEX, HIGHBORNE CAY, BAHAMAS BY R. PAMELA REID', IAN G. MACINTYRE’ and ROBERT S. STENECK? ABSTRACT Microbial deposition plays an important role in the construction of a present-day fringing reef complex at Highborne Cay, Bahamas. This reef consists of an algal ridge that grades shoreward to a back reef lagoon with a diversity of microbial buildups. Intertidal stromatolites and thrombolites form tabular mounds several meters in diameter and up to a meter thick in the nearshore zone. Shallow subtidal stromatolites form ridges and columnar heads up to half a meter high in the sandy lagoon. A tufa-like rock forms ridges at the bases of some of the thrombolites. The algal ridge is composed mainly of the branching crustose coralline alga, Neogoniolithon strictum, which, until recently, was not known to form algal ridges. The coralline algae are commonly coated with micritic crusts of possible microbial origin. This unusual reef is an ideal system for geomicrobiological studies. INTRODUCTION Laminated microbial deposits known as stromatolites have a unrivalled geologic history dating back more than 3 billion years and ranging to the present day. Modern stromatolites forming in waters of normal marine salinity were first discovered in the Schooner Cays on the northeastern margin of Exuma Sound (Dravis 1983). Subsequently, marine stromatolites have been mapped at numerous locations throughout the Exuma Cays, on the western margin of Exuma Sound (Dill et al. 1986; Dill 1991; Reid and Brown 1991; Reid et al. 1995). Stromatolites in most of these examples occur as columnar buildups in tidal channels. However, at two locations, Stocking Island and Highborne Cay, stromatolites form as part of a fringing reef complex. In addition to stromatolites, these fringing reefs incorporate an unusual algal ridge system (Steneck et al. 1997). The growth history of the stromatolite-algal ridge reef complex at Stocking Island was studied by Macintyre et al. (1996). The present paper describes the reef zonation at Highborne Cay. The Highborne Cay reef includes some of the best laminated stromatolites in the ‘Rosenstiel School of Marine and Atmospheric Science, University of Miami, 4600 Rickenbacker Cswy, Miami, FL 33149 and * Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, Washington D.C. 20560. * Department of Oceanography, University of Maine, Darling Marine Center, Walpole, Maine 04573. Manuscript received 7 July 1999; revised 27 July 1999 FIGURE 2. Fringing reef along the east shore of Highborne Cay; line marks the location of the transect shown in Figure 3. HIGHBORNE REEF PROFILE ' ALGAL A BACK REEF RIDGE A e Schizothrix mat 3 50 and Gracilaria < Schizothrix Knobby turf Neogoniolithon heads Schizothrix Schizothrix mat Smooth base es Coralline algal Is. EF] Sand 0 10 20 30 40 50 Horizontal distance (m) FIGURE 3. Transect across the fringing reef, along line A-A' as shown in Figure 2; vertical lines mark the location of probe holes made to estimate depth to Pleistocene. 3 Exuma Cays. Consequently, this island was selected as the primary field site for a multidisciplinary investigation of geomicrobiological processes forming lithified micritic laminae in modern marine stromatolites. This project, which involves participation of approximately 12 investigators from ten institutions in the US and Europe, is known as the Research Initiative on Bahamian Stromatolites, or RIBS project. In addition to stromatolites, a variety of other microbial deposits are abundant in the Highborne Cay reef, including thrombolites, micritic crusts and tufa-like deposits; these microbialites have not been studied in detail and await investigation. Highborne Cay is a small island near the north end of the Exuma Cays (Fig. 1). Formed predominantly by Pleistocene aeolianite limestone, the island consists of two curvilinear ridges, 2 to 4 km long and trending north-south; the two ridges coalesce in the middle of the island, giving a maximum width of about | km. A fringing reef extends along a beach on the eastern margin of the island, facing Exuma Sound. A shelf, with an average depth of 10-20 m, extends 1-2 km offshore; at the shelf margin water depths increase rapidly to greater than 1000 m in Exuma Sound. Surface waters of Exuma Sound have a salinity of 36-37 ppt and are saturated with respect to both aragonite and calcite (Droxler et al. 1988). Intermittent south and southeastern tradewinds dominate this area (Adey 1978); large ocean swells are rare. Tides are diurnal and have an average range of about | m. REEF DESCRIPTION The fringing reef along the eastern margin of Highborne Cay is about 2.5 km long. It is best developed in the southernmost kilometer of the beach, where it is 30 to 50 m wide, has 0.5 to 1 m of relief, and shows distinct zonation: an algal ridge colonized dominantly by macroalgae grades shoreward to a back reef lagoon of microbial buildups (Fig. 2). A transect across this southern section, measured perpendicular to the shore using a level line and probe, is shown in Fig. 3. Further north along the beach, the back reef zone of microbial buildups is lacking and the reef becomes a discontinuous algal ridge. Back Reef The back reef zone is comprised of carbonate mounds in a sandy lagoon. Surfaces of these mounds are colonized by two distinct types of microbial mat: smooth mats (Fig. 4A) and knobby turf (Fig. 4B); carbonate sand grains are abundant in both mat types. The biologic composition of the smooth mats will be described in detail by Prufert-Bebout et al. (in prep.); these mats are dominated by the filamentous cyanobacterium Schizothrix sp., but also include a variety other cyanobacteria, bacteria and microalgae, such as Oscillatoria sp., Microcoleus sp., Phormidium sp., an endolithic coccoid cyanobacterium Solentia sp., and diatoms. The knobby turf is composed dominantly of tufts of the cyanobacterium Dicothrix (Figs. 4B), but also includes microalgae, such as Cladophoropsis sp. and Ernodesmus sp., as well as macroalgae, such as Batophora sp. Diatoms and Batophora show seasonal variability, with higher numbers occurring in warm summer months. Likewise the prokaryotic mat community shows some _ seasonal FIGURE 4. (A) Smooth Schizothrix mat on the surface of an intertidal stromatolite; hammer head on left is 12 cm long. (B) Knobby turf on the surface of an intertidal thrombolite, or cauliflower head. (C) Light colored Schizotrhix mat on intertidal stromatolites (S) in foreground grades to knobby turf on thrombolites (T). (D) The same view as in (C), but beach sand is covering the nearshore stromatolites. 5) variability (Prufert-Bebout et al. in prep.). Schizothrix mats are dominant on surfaces that are frequently buried and uncovered by shifting sand; knobby turf dominates exposed mounds that are rarely, if ever, buried (Figs 4C, 4D). Sawed sections through the back reef mounds reveal two distinct types of internal structure: some buildups have distinct layering (Fig. 5A), whereas others have clotted irregular fabrics, with abundant borings made by molluscs, worms and sponges (Fig. 5B). The layered deposits are typically colonized by Schizothrix mats, whereas the unlayered deposits have a surface mat of knobby turf. Similarities between the textures and microstructures of the mounds and the overlying mats, as outlined below, argues that the mounds were constructed by communities equivalent to the present-day mats. Thus, the layered deposits are termed stromatolites, and the mounds with clotted fabrics are designated thrombolites. Because of their knobby appearance, the thrombolites are also referred to as “cauliflower” heads. Highborne Cay stromatolites and the overlying Schizothrix mats are composed primarily of well-sorted fine sand (125-250 tm), consisting mainly of peloidal grains with superficial oolitic coatings. Grains in surface mats are bound by cyanobacterial filaments; these filaments are, however, typically not calcified and are not preserved at depth. Lamination in the stromatolites is defined by differential lithification: layers are most visible in sawed sections, where lithified horizons, typically 0.5 to 1 mm thick, stand out in relief on cut surfaces (Fig 5A). Lamination in surface mats is reflected in variations in color and hardness: crusty to hard gray/green layers with high biomass alternate with soft white grain layers with low biomass (Fig. 5C). The gray/green layers contain an abundance of the endolithic cyanobacterium, Solentia sp. (Macintyre et al. in prep.; Prufert-Bebout et al. in prep.). In thin section, indurated layers in the stromatolites and Schizothrix mats appear as micritic horizons with characteristic features (Fig. 5E): they consist of a micritic crust, typically 20-40 um thick, which commonly overlies a layer of micritized carbonate grains, 200-1000 um thick. The micritized grains are often truncated along their upper surfaces and are cemented at point contacts. With increasing micritization, grain boundaries become indistinct and grains appear welded together. Size distributions of grains within the stromatolites were compared with bottom sediment from the back reef lagoon using a Coulter particle size analyzer (Model LS200). Three samples of poorly indurated stromatolite were disaggregated by soaking for 24 hours in 5% sodium hypochlorite to remove organics. Grains in these stromatolites and three samples of lagoonal sand (Fig. 6) show distinct differences in size. The stromatolites have mean grain sizes of 160-175 um, and are significantly finer than the lagoonal sediments, which have mean grain sizes of 300-350 um. In addition, the stromatolites contain 1-2 % material coarser than 400 tm, considerably less than found in the lagoonal sand (12-15 %). In contrast, the fine fraction (< 100 um) forms an average of 10 % of the stromatolites, but is insignificant in the lagoonal sediment. These differences indicate that Schizothrix mats selectively trap and bind fine sand; they exclude grains larger than 400 um, but include fine material from the water column that is winnowed out of the lagoonal sediment. In addition to trapped sediment, the fine fraction in the stromatolites includes 6 micrite precipitated within the stromatolites. At periodic intervals, stromatolite-forming Schizothrix mats may be colonized by turf or macroalgae, such as Batophora or Gracilaria. Episodes of eukaryotic colonization are evident in cross sections through the stromatolites, as decayed roots and holdfasts leave horizons of open holes. Sediment within these eukaryotic horizons is typically less well sorted, coarser grained, and more porous than that accreted by Schizothrix mats. In contrast to the laminated stromatolites, Highborne Cay thrombolites and the overlying knobby turf have an irregular microstructure (Figs 5B, 5D). The knobby surface mats are comprised of radiating tufts of filaments with entrapped carbonate sand (Fig. 5D). Thin sections show that the mats and thrombolites consist of calcified filaments (10 to 100 um in diameter), abundant micrite and carbonate grains (Fig. SF). Size distributions of sediment in these mounds have not been analyzed, but the grains appear less well sorted and coarser (medium sand-size) than in the stromatolites. Stromatolites and thrombolites at Highborne Cay form buildups of a variety of shapes and sizes (Fig. 7). Stromatolites occur as both intertidal and subtidal mounds. Intertidal stromatolites form in the nearshore beach zone (Fig. 3) where, as mentioned above, they are alternately buried and uncovered (Figs. 4C, 4D). These stromatolites grade seaward to intertidal thrombolites, which are rarely, if ever, buried (Figs. 4C, 4D). The thrombolites are prominent features of the Highborne Cay reef: they are about a meter in height and are subaerially exposed for 3 to 4 hours twice daily at low tide. Commonly protruding from beach sand at the water’s edge at low tide, the thrombolites form large, tabular mounds, that may be several meters in diameter (Figs. 3, 7A); these large mounds are typically constructed by the coalescence of smaller mounds. Thrombolites also form intertidal columns in the lagoon between the algal ridge and the nearshore tabular mounds (Figs. 3, 7A). Subtidal stromatolites are also present in this lagoon, commonly forming ridges, 20-30 cm in height, perpendicular to the shoreline (Fig. 7B). In other parts of the lagoon, subtidal stromatolites form individual heads and FIGURE 5. Internal structures of microbial buildups and surface mats in the back reef lagoon. (A) Cut section through a well laminated stromatolite-- lithified layers about 1 mm thick stand out in relief. Sample 8/98NS8TRd. (B) Cut section through a thrombolite with an irregular, clotted structure-- open bore holes are abundant. Sample SI-92-c. (C) Cross section though a Schizothrix mat; the gray/green layer at 2-3 mm depth is hard; white grain layers above and below are soft. Sample 8/98NS8TRc. (D) Cross section though knobby turf composed of Dicothrix filaments and carbonate grains-- crusty patches are dispersed throughout the knob. Sample 8/98NSc. (E) Thin section photomicrograph showing the characteristic microstructure of lithified laminae in stromatolites and Schizothrix mats: a micritic crust (c) overlies a layer of micritized sand grains (arrow), which are welded together at point contacts. Thin section 6/97NS8f. (F) Thin section photomicrograph showing the characteristic microstructure of thrombolites and knobby turf: calcified filaments (white arrow) and irregularly dispersed micritic precipitates (black arrow) are abundant. Thin section 8/98NSc. Volume % Volume % Vol. % > 400 um Differential Volume (Average) Stromatolite Sample 1 Vol. % < 100 um = 10.10 Vol. % > 400 um = 2.42 Mean = 173 um Particle Diameter (um) Differential Volume (Average) | Sediment Sample 1 0 12.89 Mean = 305 um Vol. % <100 um 1 10 100 Particle Diameter (ym) 1000 Volume % Volume % | 10-4 Differential Volume (Average) Stromatolite Sample 2 Vol. % <100um = 10.64 Vol. % > 400 um = 1.98 Mean = 161 um Particle Diameter (um) Differential Volume (Average) Sediment Sample 2 Vol. % < 100 um 0 Vol. % > 400 um = 11.72 Mean = 311 um 1 10 100 1000 Particle Diameter (um) Volume % Volume % Differential Volume (Average) 'S Stromatolite 12 Sample 3 . |Vol.% <100um = 10.32 114 Vol. % > 400 um = 0.54 Mean = 160 um Particle Diameter (um) Differential Volume (Average) 114 Sample 3 Vol. % <100um = 0.01 104 Vol. % > 400 um = 14.93 Mean = 341 um Particle Diameter (um) FIGURE 6. Histograms comparing grain-size distribution patterns in three subtidal stromatolites and three sediment samples from the back-reef lagoon. ) columns, up to 0.5 m in height (Fig. 7C). At mean low tide, the subtidal stromatolites are submerged in 30-50 cm of water. Like the intertidal stromatolites in the beach zone, the subtidal stromatolites in the lagoon are continually buried and uncovered by migrating sand (Figs 7C, 7D). The bases of the back reef thrombolites and stromatolites are generally buried and consequently little is known about the substrates on which these mounds are growing. We did however, recover what appears to be a tufa forming ridges several centimeters high near the bases of some thrombolites. This ridge rock is extremely porous (Figs. 8A). Thin sections show that carbonate sand grains are rare and that the rock is mostly micrite; in places, the micrite forms knobby projections and vague lamination (Fig. 8B). The micrite has an unusual and distinct petrographic appearance: in plane polarized light, it appears as an intimate mixture of golden and dark brown crystals that form aggregates and irregular clustered masses within a porous network (Fig. 8C). These clusters of golden and dark brown crystals are commonly rimmed with acicular cement (Fig. 8C). Scanning electron microscope (SEM) observations show that the rock consists mainly of filament molds, 3 to 10 um in diameter, encased in spherulitic clusters of aragonite needles (Figs 8D, 8E). It is therefore tentatively classified as a tufa. Formation of tufa in a subtidal marine environment would, however, be highly unusual, as tufa is typically considered to be a freshwater deposit (Pentecost and Riding 1986). The origin of this ridge rock thus warrants further investigation. Algal Ridge An algal ridge, which forms a platform about 10 m wide, forms the seaward margin of the Highborne reef and is the focus of breaking waves (Figs. 3, 9A). Hemispherical heads of the branching crustose coralline alga Neogoniolithon strictum dominate the seaward edges of the ridge, forming an emergent lip that is subaerially exposed at extreme low tides (Fig. 9B). These heads, which can be over 50 cm in diameter and 20 cm in height, are among the largest observed for N. strictum in the tropical Atlantic and Caribbean. Several species of macroalgae attach to, or become entangled in, the N. strictum heads, including Sargassum and Gracilaria. Gracilaria also forms thickets shoreward of N. strictum, grading into Schizothrix mats at the edge of the platform (Figs. 3, 9C). Patches of Schizothrix mat are also present on the seaward lip of the ridge, growing between heads, and sometimes coating, N. strictum. The sea urchin, Echinometra lucunter has a patchy distribution on the ridge, with an average density of 6 urchins/m”; this is considerably lower than the average density of 55 urchins/m? on the algal ridge at Stocking Island (Steneck et al.1998). In the subsurface, the algal ridge consists mainly of dense, well indurated coralline algal limestone. Branches of N. strictum are abundant (Figs. 10A, 10C). In addition, N. strictum and other coralline algae form platy encrustations, which are commonly intergrown with the foraminifer, Homotrema rubrum (Figs. 10B, 10D). Branches of N. strictum are often coated with micritic crusts, which may be up to several millimeters thick (Figs 10C, 10E). In thin section, these micritic crusts are similar to the tufa: in plane 11 polarized light, they appear as intimate mixtures of golden and dark brown crystals forming irregular clusters in a porous network (Fig. 10OE). SEM shows that these crusts consist of micritic crystals with a variety of shapes, ranging from granules to platy crystals and needles; these crystals are permeated with holes, less than 5 um in diameter, which could be filament molds (Fig. 10F). The crusts are of possible microbial origin, but further studies are needed to confirm this interpretation. DISCUSSION AND CONCLUSIONS The microbialite/algal ridge complex at Highborne Cay is an unusual and intriguing reef complex. Reef zonation is similar to that at Stocking Island, as described by Macintyre et al.1996. At Stocking Island, an outer ridge constructed primarily by N. strictum forms a protected back reef lagoon where conditions are favorable for the development of microbial mounds. Radiocarbon dates indicate that growth of the algal ridge at Stocking Island was initiated about 4000 years ago when the Pleistocene terrace was flooded during the Holocene transgression (Macintyre et al.1996). Maximum reef development at Stocking Is. occurred about 1500 years ago, with the development of an emergent coralline algal lip. For the past 500 years, the Stocking Is. reef has been in a destructive stage, with extensive bioerosion by E. lucunter degrading the algal ridge. As E. lucunter is considerably less abundant at Highborne Cay (6 urchins/m?) than Stocking Island (55 urchins/m?), bioerosion rates are lower and the Highborne Cay algal ridge has maintained an emergent lip. Prior to the discovery of the Stocking Island and Highborne Cay reefs (Steneck et al.1997), algal ridges built by N. strictum were unknown. Algal ridges are typically composed of robust coralline algae, which form thick laminar layers that can withstand high wave energy (e.g. Adey 1975; Bosence 1984). Indeed, high wave energy has been considered essential for algal ridge formation, as waves deter parrotfish, which are major herbivores of coralline algae. (Adey 1975 1978; Steneck and Adey 1976). The Exuma algal ridges are unique in many respects. They are formed by a relatively unusual coralline species, N. strictum, which had not been reported to build algal ridges. N. strictum thrives in environments with relatively low wave energy, high rates of sedimentation and low rates of herbivory, where most ridge-forming algae cannot grow (Steneck et al.1997). Experimental manipulations at Stocking Island, which showed that N. strictum is capable of surviving when covered with sediment for at least 100 days, indicate that this alga is adapted to surviving in sediment-dominated environments (Steneck et al. 1997). FIGURE 7. Microbial buildups in the back reef lagoon. (A) View looking north along the beach. Intertidal thrombolites (T) form tabular mounds in the beach zone and isolated heads in the lagoon, shoreward of the algal ridge (AR). Subtidal stromatolites (S) are present in the foreground; stakes mark the position of study sites for the RIBS project. (B) Ridges of subtidal stromatolites, which trend perpendicular to the shore. (C) Subtidal columnar stromatolites. (D) The same stromatolites as shown in C, but a few days later when the heads are almost entirely covered by sand. 12 Inherently low levels of parrotfish grazing in reefs along the west margin of Exuma Sound may be important in allowing algal ridge development in this moderate wave energy environment (Steneck et al. 1997). Growth of stromatolites in the lee of an algal ridge is another unique feature of the Highborne Cay and Stocking Island reefs. As the first known examples of marine stromatolites in a modern reef environment, these structures provide an opportunity to examine ecological controls of stromatolite development. Manipulative experiments conducted by Steneck et al. (1998) showed that stromatolites dominate where species diversity (especially among eukaryotes) and associated ecologic pressures is low. At Highborne Cay and Stocking Is, frequent periods of sediment inundation in the back reef create an ecologic refuge for stromatolites (Steneck et al. 1998). Most eukaryotic reef- dwelling organisms cannot survive or colonize under high sedimentation rates. Turf algae, for example, can survive desiccation and thermal stress on a reef flat, but not the high sedimentation rates of the stromatolite zone (Steneck et al. 1998). In contrast, stromatolite-forming Schizothrix mats thrive under conditions of high sedimentation. Although these mats can also grow in other areas of the reef environment, their reef- building contribution is low and their laminated microstructure is lost except in habitats where abiotic stresses maintain an ecologic refuge-- i.e. a refuge from the ecologic pressures of other organisms (Steneck et al. 1998). As recognized by Reid et al. (1995) and Macintyre et al. (1996), lamination in Exuma stromatolites results from periodic formation of lithified micritic laminae within cyanobacterial mats. Introduction of turf algae disrupts the continuity of the flat-lying, stromatolite-forming community and inhibits the formation of laminae, resulting in formation of thrombolites rather than stromatolites (Reid et al.1995; and Macintyre et al. 1996). At Highborne Cay, stromatolites form in the beach zone and in lagoonal areas where they are continually buried and uncovered by migrating sand. Stromatolites grade into thrombolites where sedimentation rates decrease sufficiently to allow growth of Dicothrix, Cladophoropsis and other turf algae. The microbial mats forming Highborne Cay stromatolites are composed almost exclusively of prokaryotic communities (Prufert-Bebout et al. 1999; Pinckney et al. 1995; Pinkney and Reid 1997) and exhibit superb lamination. Ongoing studies in the RIBS project are documenting biogeochemical processes of lithification within these mats. For example, Macintyre et al. (in review) showed that periodic introduction of the coccoid endolith Solentia sp. to the Schizothrix community during episodes of low sedimentation FIGURE 8. Tufa-like rock at the bases of some thrombolites; Sample SI-92-132. (A) Hand specimen showing the porous, knobby nature of this rock. (B, C) Thin section photomicrographs in plane polarized light showing that the rock is composed mainly of micrite with pores of varying sizes and shapes. High magnification view in (C) shows that the micrite (M) forms irregularly-shaped clusters of golden and dark brown crystals; these clusters are fringed with acicular cement (arrows) precipitated in open pore space (P). (D, E) SEM photomicrographs showing that the rock is composed almost entirely of filament molds encased in aragonite needles. 14 results in formation of lithified layers of micritized grains. In addition, sulfur cycling within the mats is discussed by Visscher at al (1998), who showed that photosynthesis coupled to sulfate reduction and sulfide oxidation is more important than photosynthesis coupled to aerobic respiration in stromatolite lithification. Additional papers documenting stromatolite microstructure, nitrogen cycling, polymer production and degradation etc. are currently in preparation. In contrast to the intensive studies of Highborne Cay stromatolites, there have been no detailed investigations of the other microbial deposits within this unusual reef system-- the cauliflower heads, or thrombolites, which form prominent intertidal back reef deposits; the ridges of tufa-like material at the bases of the thrombolites; and the micritic crusts coating coralline algae in the algal ridge. Highborne Cay is indeed a haven for microbial deposition and a natural laboratory that invites further geomicrobiological investigations. ACKNOWLEDGEMENTS This research was supported by National Science Foundation Grants OCE- 9116296 to R.P. Reid and R.S. Steneck and OCE-9530215 to R.P. Reid. Thanks to D.A. Dean for thin section preparation; W.T. Boykins for size analysis and drafting; and the management of Highborne Cay and crew of the R/V Calanus for logistical assistance at the field site. This paper is a contribution to IGCP Project 380, Biosedimentology of Microbial Buildups. RIBS Contribution Number 4. FIGURE 9. Algal ridge at the seaward edge of the Highborne reef. (A) View looking south across the ridge. Coralline algae limestone forms a platform about 10 m wide; the branched coralline alga, N. strictum (N), forms an emergent lip at the seaward edge of this platform. (B) Closer view of N. strictum at the platform edge. (C) Gracilaria (G) grading shoreward to Schizothrix mats (S) at the edge of the ridge. REFERENCES Adey, W.H. (1975). The algal ridges and coral reefs of St. Croix: their structure and Holocene development. Atoll Research Bulletin 187:1-67. Adey, W.H. (1978). Algal ridges of the Caribbean sea and West Indies. Phycologia 17:361-367. Bosence, D.W.J. (1984). Construction and preservation of two modern coralline algal reefs, St. Croix, Caribbean. Paleontology 27:549-574. Dill, R.F. (1991). Subtidal stromatolites, ooids and crusted-lime muds at the Great Bahama Bank Margin. Jn Osborne, R.H. (ed.), From Shoreline to Abyss, SEPM Special Publication 46:147-171, Tulsa. Dill, R.F., Shinn, E.A., Jones, A.T., Kelly, K., and Steinen, R.P. (1986). Giant subtidal stromatolites forming in normal salinity water. Nature 324:55-58. Dravis, J.J. (1983) Hardened subtidal stromatolites -Bahamas. Science, 219:385-386. Droxler, A.W., Morse, J.W., and Kornicker W.A., 1988, Controls on carbonate mineral accumulation in Bahamian basins and adjacent Atlantic Ocean sediments: Journal of Sedimentary Petrology 58:120-130. Macintyre, I.G., Reid, R.P., and Steneck, R.S. (1996). Growth history of stromatolites in a fringing Holocene reef, Stocking Island, Bahamas. Journal of Sedimentary Research 66:231-242. Pentecost, A. and Riding, R. (1986). Calcification in cyanobacteria in Leadbeater, B.S.C. and Riding, R. (eds.) Biomineralization in Lower Plants and Animals. Systematics Association Special Volume No. 30, Clarendon Press, Oxford, p. 73-90. FIGURE 10. Samples from the algal ridge. (A, B) Hand specimens showing WN. strictum in branched (A, arrows) and encrusting (B, arrows) forms; algal crusts in B are intergrown with the foraminifer H. rubrum. (C, D) Thin section photomicrographs in plane polarized light showing the microstructure of branched and encrusting N. strictum (N); arrow in C indicates micritic coatings on branches of N. strictum. (C, thin section SI-92-56; D, thin section SI-93-125). (E) Higher magnification photomicrograph showing a more detailed view of a micritic crust (arrow) on a branch of N. strictum (N); the micrite is an intimate mixture of golden and dark brown crystals with high porosity. Thin section SI-92-56. (F) SEM photomicrograph of a micritic crust; granular, platy and needle shaped crystals are permeated by holes, 1-3 um in diameter, which may be filament molds Sample SI-92-55. 18 Pinckney, J., Paerl, H.W., Reid, R.P. and Bebout, B. (1995). Ecophysiology of stromatolitic mats, Stocking Island, Exuma Cays, Bahamas. Microbial Ecology, v. 29, p. 19-37. Pinckney, J.L. and Reid, R.P. (1997). Productivity and community composition of stromatolitic microbial mats in the Exuma Cays, Bahamas. Facies 36:204-207. Reid, R. Pamela and Browne K.M. (1991). Intertidal stromatolites in a fringing Holocene reef complex, Bahamas. Geology 19:15-18. Reid, R.P., Macintyre, I.G., Steneck, R.S., Browne, K.M., and Miller, T.E. (1995). Stromatolites in the Exuma Cays, Bahamas: Uncommonly common. Facies 33:1-18. Steneck, R.S. and Adey, W.H. (1976). The role of environment in control of morphology in Lithophyllum compressum, a Caribbean algal ridge builder. Botanica Marina 19:197- DiS: Steneck, R.S., Macintyre, I.G. and Reid, R.P. (1997). Unique algal ridge systems of Exuma Cays, Bahamas. Coral Reefs 16:29-37. Steneck, R. S., Miller, T. E., Reid, R. P. and Macintyre, I. G. (1998). Ecological controls on stromatolite development in a modern reef environment: a test of the ecological refuge paradigm. Carbonates and Evaporites 13:48-65. Visscher, P.T., Reid, R.P., Bebout, B.M., Hoeft, S.E., Macintyre, I.G., and Thompson, J. Jr. (1998). Formation of lithified micritic laminae in modern marine stromatolites (Bahamas): the role of sulfur cycling. American Mineralogist 83:1482-1491. ATOLL RESEARCH BULLETIN NOS. 459-465 NO. 459. NO. 460. NO. 461. NO. 462. NO. 463. NO. 464. NO. 465. SPECIES RICHNESS OF RECENT SCLERACTINIA BY STEPHEN D. CAIRNS ATOLLS AS SETTLEMENT LANDSCAPES: UJAE, MARSHALL ISLANDS BY MARSHALL I. WEISLER REPORT ON FISH COLLECTIONS FROM THE PITCAIRN ISLANDS BY JOHN E. RANDALL FISH NAMES IN LANGUAGES OF TONGA AND FIJI BY R. CHRISTOPHER MORGAN THE NON-NATIVE VASCULAR PLANTS OF HENDERSON ISLAND, SOUTH CENTRAL PACIFIC OCEAN BY STEVE WALDREN, MARSHALL I. WEISLER, JON G. HATHER AND DYLAN MORROW REVISED VEGETATION CLASSIFICATION OF TURNEFFE ATOLL, BELIZE BY MALCOLM R. MURRAY, SIMON A. ZISMAN AND CHRISTOPHER D. MINTY A MICROBIALITE/ALGAL RIDGE FRINGING REEF COMPLEX, HIGHBORNE CAY, BAHAMAS BY R. PAMELA REID, IAN G. MACINTYRE AND ROBERT S. STENECK ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. AUGUST 1999 ATOLL RESEARCH BULLETIN NOS. 466-480 NATURAL HISTORY OF THE PELICAN CAYS, BELIZE EDITED BY RESEARCH IAN G. MACINTYRE AND KLAUS RUTZLER BULLETIN Issued by NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C. U.S.A. MARCH 2000 ATOLL RESEARCH BULLETIN NO. 466-480 NATURAL HISTORY OF THE PELICAN CAYS, BELIZE EDITED BY IAN G. MACINTYRE AND KLAUS RUTZLER ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 The Atoll Research Bulletin is issued by the Smithsonian Institution to provide an outlet for information on the biota of tropical islands and reefs and on the environment that supports the biota. The Bulletin is supported by the National Museum of Natural History and is produced by the Smithsonian Press. This issue is partly financed and distributed with funds from Atoll Research Bulletin readers and authors. The Bulletin was founded in 1951 and the first 117 numbers were issued by the Pacific Science Board, National Academy of Sciences, with financial support from the Office of Naval Research. Its pages were devoted largely to reports resulting from the Pacific Science Board's Coral Atoll Program. All statements made in papers published in the Atoll Research Bulletin are the sole responsibility of the authors and do not necessarily represent the views of the Smithsonian nor of the editors of the Bulletin. Articles submitted for publication in the Atoll Research Bulletin should be original papers in a format similar to that found in recent issues of the Bulletin. First drafts of manuscripts should be typewritten double spaced and can be sent to any of the editors. After the manuscript has been reviewed and accepted, the author will be provided with a page format with which to prepare a single-spaced camera-ready copy of the manuscript. COORDINATING EDITOR Ian G. Macintyre National Museum of Natural History Smithsonian Institution ASSISTANTS Washington, D.C. 20560-0125 Kasandra D. Brockington William T. Boykins, Jr. Kathryn Clark-Bourne EDITORIAL BOARD Zip Codes Stephen D. Cairns 20560-0163 National Museum of Natural History Brian F. Kensley 20560-0163 Smithsonian Institution Mark M. Littler 20560-0166 Washington, D.C. Wayne N. Mathis 20560-0169 Jeffrey T. Williams 20560-0159 Joshua I. Tracey, Jr. 20560-0137 Warren L. Wagner 20560-0166 Roger B. Clapp National Museum of Natural History National Biological Survey, MRC-111 Smithsonian Institution Washington, D.C. 20560-0111 David R. Stoddart 837 Oxford Street Berkeley, CA 94707-2013 Bernard M. Salvat Ecole Pratique des Hautes Etudes Labo. Biologie Marine et Malacologie Université de Perpignan 66025 Perpignan Cedex, France PUBLICATIONS MANAGER A. Alan Burchell Smithsonian Institution Press ATOLL RESEARCH BULLETIN NOS. 466-480 PAGES NO. 466. | ORIGIN OF THE PELICAN CAYS PONDS, BELIZE 1 BY IAN G. MACINTYRE, WILLIAM F. PRECHT, AND RICHARD B. ARONSON NO. 467. A GENERAL BIOLOGICAL AND GEOLOGICAL SURVEY OF THE RIMS OF PONDS IN THE 15 MAJOR MANGROVE ISLANDS OF THE PELICAN CAYS, BELIZE BY IAN G. MACINTYRE, IVAN GOODBODY, KLAUS RUTZLER, DIANE S. LITTLER, AND MARK M. LITTLER NO. 468. | MANGROVE PEAT ANALYSIS AND RECONSTRUCTION OF VEGETATION HISTORY AT THE 47 PELICAN CAYS, BELIZE BY KAREN L. MCKEE AND PATRICIA L. FAULKNER NO. 469. | PRELIMINARY HYDROGRAPHIC SURVEYS OF SOME PONDS IN THE PELICAN CAYS, BELIZE 61 BY DANIEL W. URISH NO. 470. | HYDROGRAPHY OF A SEMI-ENCLOSED MANGROVE LAGOON, MANATEE CAY, BELIZE 91 BY TRACY A. VILLAREAL, STEVE L. MORTON, AND GEORGE B. GARDNER NO. 471. | COMMUNITY STRUCTURE, WATER COLUMN NUTRIENTS, AND WATER FLOW IN TWO PELICAN 107 CAYS PONDS, BELIZE BY THOMAS A. SHYKA AND KENNETH P. SEBENS NO. 472. PHYTOPLANKTON ECOLOGY AND DISTRIBUTION AT MANATEE CAY, PELICAN CAYS, BELIZE 125 BY STEVE L. MORTON NO. 473. DINOFLAGELLATE ASSOCIATIONS IN A CORAL REEF-MANGROVE ECOSYSTEM: PELICAN AND 135 ASSOCIATED CAYS, BELIZE BY MARIA A. FAUST NO. 474. | CHECKLIST OF MARINE ALGAE AND SEAGRASSES FROM THE PONDS OF THE PELICAN 153 CAYS, BELIZE BY DIANE S. LITTLER, MARK M. LITTLER, AND BARRETT L. BROOKS NO. 475. | EPIPHYTIC FORAMINIFERA OF THE PELICAN CAYS, BELIZE: DIVERSITY AND DISTRIBUTION 209 BY SUSAN L. RICHARDSON NO. 476. | DIVERSITY OF SPONGE FAUNA IN MANGROVE PONDS, PELICAN CAYS, BELIZE 231 BY KLAUS RUTZLER, MARIA CHRISTINA DIAZ, ROB W.M.VAN SOEST, SVEN ZEA, KATHLEEN P. SMITH, BELINDA ALVAREZ, AND JANIE WULFF NO. 477. SPONGE PREDATORS MAY DETERMINE DIFFERENCES IN SPONGE FAUNA BETWEEN TWO SETS 251 OF MANGROVE CAYS, BELIZE BARRIER REEF BY JANIE L. WULFF NO. 478. | GNATHOSTOMULIDA IN THE PELICAN CAYS, BELIZE 267 BY WOLFGANG STERRER NO. 479. | ECHINODERMS OF THE RHOMBOIDAL CAYS, BELIZE: BIODIVERSITY, DISTRIBUTION, AND 275 ECOLOGY BY GORDON HENDLER AND DAVID L. PAWSON NO. 480. | DIVERSITY AND DISTRIBUTION OF ASCIDIANS (TUNICATA) IN THE PELICAN CAYS, BELIZE 303 BY IVAN GOODBODY ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 Aerial views of the Pelican cays. Top: Looking northeast across Co Cat Cay in foreground, with Manatee Cay, Fisherman’s Cay, Ridge Cay, Northeast Cay (right to left) in the middle ground. Other rhomboid cays in background. Bottom Left: Detail of ridge and pond system in Cat Cay. Bottom Right: Overview of ponds in Manatee Cay (foreground) and Fisherman’s Cay. PREFACE With the continuing decline in the world’s natural resources, a question of prime public concern is how to assess the health of biological communities and ecosystems. As biologists have long emphasized, species richness is the best indicator of environmental quality, but it cannot be evaluated without substantial knowledge of the animals and plants inhabiting the Earth. That is particularly the case for the planet's oceans, although the marine sciences have amassed considerable information over the past five decades through advances in technology. Mask and snorkel, scuba diving, sophisticated research submersibles, and a variety of other equipment have enabled researchers to make direct and detailed observations, collect samples, study the interactions of species with one another and with their environment, and conduct experiments in situ. Still, biodiversity and interactive processes in the sea are much less visible and less well understood than those in terrestrial systems; hence the public remains ill equipped to protect coastal environments and manage marine resources. The dimensions of the problem are particularly evident in the Caribbean region, where population pressures, poor data on environmental quality, limited protective legislation or enforcement, and recent natural disasters such as hurricanes and solar-radiation stress have all contributed to the degradation of reefs, mangroves, seagrass meadows, and other important coastal communities. In Belize, however, a small human population, cautious economic development, and a late-arriving tourist boom have left many of the spectacular marine ecosystems intact. More than 200 km of the "Meso-American" barrier reef crowned by coral islands parallel the mainland of Belize and shelter an extensive lagoon with patch reefs and hundreds of mangrove cays. The lagoon is up to 25 km wide and an important fisheries resource for the country. More than 25 years ago, scientists of the Smithsonian Institution’s National Museum of Natural History established a marine field station on tiny Carrie Bow Cay, one of Belize's barrier-reef coral islands, to study the diversity and ecology of marine communities. This initiative grew into a comprehensive, long-term research program named Caribbean Coral Reef Ecosystems Program (CCRE). Since 1994, when Paul and Mary Shave guided CCRE scientists to a poorly charted, virtually unknown group of mangrove islands called the Pelican Cays in the southern Belize lagoon at approximately 16°39.8' N, 88°11.5' W, CCRE has given a great deal of attention to this area. The Pelican Cays resemble thousands of other mangrove islands throughout the Caribbean but with several intriguing differences. Here, red mangrove is anchored on top of a live and lush coral reef, not in mud, as is the case elsewhere in the lagoon, and several of the cays exhibit unique physical characteristics such as deep blue lagoon-like ponds encircled by steep, lush coral ridges. CCRE's initial reconnaissance of the Pelican Cays archipelago uncovered species richness and live surface cover that are unparalleled in the Caribbean. The reef, mangrove-root, and peat substrates are thickly overgrown by several layers of brilliantly colored organisms, including sponges, ascidians, seaweeds, and corals. Although we do not yet quite understand the causes of this unusually high biological diversity, we fear for its future. Because the coral substrate is fragile and fine calcareous sediments are abundant, it would take just a few careless visitors over a short period of time to set in motion breakage and siltation that could irreparably harm these delicate communities. Recent water warming and extended periods of calm are already taking their toll, as is shown in the following pages. Recognizing the importance of preserving marine resources, the government of Belize has begun setting up several marine parks and protected zones, which include the Pelican Cays. This collection of papers contributes toward that end by providing a biodiversity assessment, geological and paleoecological data, and an evaluation of the environmental parameters that maintain the balance of the unusual ecosystem called Pelican Cays. Ian G. Macintyre Klaus Riitzler Editors Washington, D.C. November, 1999 ACKNOWLEDGMENTS — In Belize we thank Paul and Mary Shave for introducing us to the magnificent world of the Pelican Cays, staff of the Belize Coastal Zone Management and Fisheries Unit for encouraging our studies, and Tony Rath for contributing photographs. Several colleagues at the Smithsonian Institution, Washington, D.C., helped with the production of this volume, particularly Molly K. Ryan (maps and graphic design) and William T. Boykins and Kathleen P. Smith (editorial and electronic processing). Venka V. Macintyre, Falls Church, VA, edited the final manuscripts. ATOLL RESEARCH BULLETIN NO. 466 ORIGIN OF THE PELICAN CAYS PONDS, BELIZE BY IAN G. MACINTYRE, WILLIAM F. PRECHT, AND RICHARD B. ARONSON ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 Carbonate Shoals (<2m depth) Zz Mangrove Cays Caribbean Sea Guatemala Honduras / Manatee Cay o A Cat Cay PELICAN ° CAYS Figure 1. Index map showing the location of the Pelican Cays in the Belizean Barrier Reef Complex. Modified from a Landsat TM image acquired 18 September 1987. ORIGIN OF THE PELICAN CAYS PONDS, BELIZE BY IAN G. MACINTYRE,' WILLIAM F. PRECHT,’” and RICHARD B. ARONSON? ABSTRACT Probing with interlocking steel rods and short cores indicates that the small ponds characteristic of the Pelican Cays are formed by differential coral accumulations on a polygonal karst pattern eroded into the underlying Pleistocene limestone. Rapidly accumulating Acropora cervicornis—dominated communities have been responsible for exaggerating the karst relief, forming steep-sided ridge patterns that commonly result in small restricted ponds. A shallowing- upward facies pattern is documented within the ridges and consists of an Acropora cervicornis facies grading into a Porites divaricata facies, which finally gives way to mangrove peat. INTRODUCTION The most striking topographic feature of the Pelican Cays (Fig. 1), in the south-central lagoon of the Belizean Barrier Reef, is the complex network of coral ridges, both submerged and exposed, some with mangrove cover. This unusual honeycomb topographic pattern, once colonized by the red mangroves Rhizophora mangle, forms the characteristic enclosed or partly enclosed ponds of this area of the barrier reef complex (Fig. 2). Similar ridge patterns have been reported in shallow lagoon areas in the Maldives by Purdy and Bertram (1993), who related them to the karst topography of the underlying limestone. The purpose of this study was to investigate the role of subsurface control in the formation of the Pelican Cays network of reef ridges, the extent to which Holocene differential reef accumulation contributes to the relief of these ridges, and the relationship of these ridge patterns to the origin of enclosed ponds. 'Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560-0125. °PBS & J Engineering Science, 2001 Northwest 107" Avenue, Miami, FL 33172. *Dauphin Island Sea Lab, 101 Bienville Boulevard, Dauphin Island, AL 36528. Figure 2. Aerial view of the Pelican Cays looking SSE: Fisherman's Cay (left), Manatee Cay (right), and Cat Cay (background). Note the characteristic enclosed and partially enclosed ponds in these islands and the network of shallow ridges in Pond C in Manatee Cay. ( Photo by T. Rath) METHODS The research focused on the well-developed network of ridges in Pond C of Manatee Cay (Fig. 2). A 0.95-cm-diameter steel probe, in 3.05-m extensions, was used to establish the contact with the underlying rock substrate (Fig. 3). This contact was easy to detect because of the relative ease of penetrating the overlying open branching coral framework consisting predominantly of Acropora cervicornis (Shinn et al., 1979; Westphall, 1986; Aronson and Precht, 1997). In addition, three short cores were collected by pushing a 7.6-cm-diameter aluminum tube into the crests of reef ridges. Core logs were expanded to correct for compaction that occurred during coring. Radiocarbon dates were determined by standard techniques by Beta Analytic Inc. (Miami, Florida). These dates are reported as radiocarbon years before A.D. 1950, conventionally termed "before present" (B.P.), using a Libby half-life of 5,568 yrs and a modern standard based on 95% of the activity of the National Bureau of Standards’ oxalic acid. No corrections were made for the DeVries effect, reservoir effect, or natural isotopic fractionation. RESULTS Our probing and coring activities were limited to the complex ridge system around the opening to the large central pond in Manatee Cay, Pond C (Fig. 4). One transect of probe sites, including a core site, was established across the coral ridge that runs across the entrance to this Figure 3. Probing open coral framework with connecting steel rods to locate depth of solid rock substrate. pond, and a similar transect was set up across the northern inner ridge that cuts across the pond. In addition, a probe and core site, was also located on the crest of the southern inner ridge, an a 3-m probe section was pushed into the center of the southern pond and did not encounter a hard base. As can be seen in Table 1, the two probe transects indicate that the ridges are established on minor relief (1 m and 0.7 m) at the edge of a significant drop in elevation of the hard rock substrate (8.6 m and 5.7 m). This is well illustrated in Fig. 5, which shows a transect across the ridge that extends across the mouth of Pond C. Although the hard substrate was not located in the southern pond probe, similar relief was encountered between this site and the site on the southern inner ridge. The accumulation of Holocene sections (Table 1) indicates that reef growth on the ridges is approximately twice that found in the interior of ponds. Core | was collected on the crest of the ridge at the mouth of Pond C and consisted of two core intervals. The first interval reached a depth of 1.5 m and had a core recovery of 80%. The second core interval was from 1.5 m to 4.1 m and had a recovery of 72%. As the core log (Fig. 6) indicates, most of the coral recovered was Acropora cervicornis, with smaller amounts of Agaricia tenuifolia, Porites furcata, and Millepora spp. There was a marked increase in the amount of A. fenuifolia and Millepora spp. in the top section of the core. This transition in reef facies is related to the colonization of A. tenuifolia following the 1986-1992 mass mortality of A. cervicornis throughout this area of the lagoon (Aronson and Precht, 1997). A light grey mud matrix was found throughout both core intervals. [e) e Mangrove (5 Coral ridges Manatee Cay rs Probe sites ¥ Probe and core sites Depth below M.S.L. Loitisii) in meters (for probe only) 100 m Figure 4. Map showing probe and core locations in the central area of Pond C, Manatee Cay. DEPTH BELOW SEALEVEL (METERS) 10 ROCK BASE 30 40 HORIZONTAL DISTANCE (METERS) Figure 5. Probe transect across ridge at mouth of Pond C, Manatee Cay. The hard rock substrate shows a slight elevation in relief before a significant increase in depth. Note the exaggeration of this rim relief by the differential accumulation of the overlying section. Table 1. Probe Holes in Pond C , Manatee Cay (depths in meters) Location Pond Entrance Ridge Base outer slope Ridge crest Base inner slope Northern Inner Ridge Base northern slope Ridge crest Base southern slope Southern Inner Ridge Ridge crest Southern Inner Pond Center of pond Water depth 0.8 14.3 Probe depth 9.0 Byler Depth of Pleistocene (below MSL) 20.9 12.3 13.3 19.0 13.3 14.0 9.8 17.4+ Core 1 Ridge at mouth of pond Water depth 0.6m 0-1.5m Mostly freshly preserved Acropora cervicornis with traces of Agaricia tenuifolia and Porites divaricata. Concentration of Millepora spp. and Agaricia tenuifolia in Halimeda sand in top 0.5m. All in a light grey mud matrix. * TC 70+60 BP 1.5-4.1m Dominantly Acropora cervicornis with traces of Agaricia tenuifolia, Millepora spp. and Porites furcata. Halimeda sand at 1.9-2.2m. All ina light grey mud matrix. DEPTH IN METERS ig) * [FC 780 + 60 BP] LEGEND Ta : ; y Acropora cervicornis 7 Branching Porites spp. (i Agaricia tenuifolia & ~~ Millepora spp. tz: Sand == Mud Figure 6. Graphic summary of data from the two core intervals of Core 1. Recovery was only 21% for Core 2, which was located on the crest of the northern inner ridge. This core (Fig. 7) showed a sharp transition in reef facies at a depth of 2.14 m. The upper section consisted of scattered Porites divaricata, with only a trace of A. cervicornis fragments in a brown organic-rich mud matrix; in contrast, the lower section is dominantly A. cervicornis, with some Agaricia sp. and Porites spp. in a light grey mud matrix. A similar transition of reef facies was also noted in Core 3 (Fig. 8), which was collected on the crest of the southern inner ridge. The upper 0.5 m consisted predominantly of P. divaricata with only traces of A. cervicornis, over a dominantly 4. cervicornis lower interval. Core 2 Northern Inner Ridge Water Depth 1.0m 0-2.14m Scattered Porites divaricata in brown organic-rich mud. A few fragments of Acropora cervicornis. . 2.14-3.8m Mostly Acropora cervicornis with some Porites furcata and a trace of Agaricia sp. Large Porites porites at base. All in a light grey mud matrix. DEPTH IN METERS ine) LEGEND p72 3 x Acropora cervicornis Vd Branching Porites spp. Sand == Mud 4 Figure 7. Graphic summary of data from Core 2. Core 3 Southern Inner Ridge Vater depth 0.8m 0) 0-0.5m Mostly Porites divaricata with some Acropora cervicornis in a grey mud matrix oO 0.5-0.8m Acropora cervicornis with some Agaricia sp. aa and Porites sp. fragments in mud grey mud matrix. ii 0.8-1.6m Acropora cervicornis in grey mud matrix Sis Zz = o Ww oO LEGEND Acropora cervicornis 1) 2 ( Branching Porites spp == Mud Figure 8. Graphic summary of data from Core 3. Coral samples from both core holes and the base of peat sections exposed in undercut edges of ponds (Fig. 9) were radiocarbon dated (Table 2). Dates ranged from 780 + 60 yrs B.P. to 70 + 60 yrs B.P. Figure 9. Undercut exposure in the edge of Pond C, Manatee Cay, showing mangrove peat overlying coral framework. Table 2. Radiocarbon Dates Location Material dated Depth below MSL From base of mangrove P. divaricata 1.33m 190 +60 undercut, east side of Pond A, Cat Cay Core from ridge at entrance A. cervicornis 2.1m 70 +60 to Pond C, Manatee Cay A. cervicornis 3.8m 780 +60 Core from northern inner A. cervicornis 3.1m 510 +80 ridge, Pond C, Manatee Cay From base of mangrove P. divaricata 1.33m 190 +60 undercut, east side of Pond A, Cat Cay From base of mangrove A. cervicornis 1.28m 590 +60 undercut, south side of Pond C, DISCUSSION The Pelican Cays are characterized by an unusual network of reef ridges that are both submerged and emergent, some with mangrove overgrowths. This network of ridges is responsible for the formation of the ponds in this area, which are the habitat of a great diversity of marine life. The distinctive feature of the network pattern is that the ridges commonly intersect at right angles. This pattern is identical to the polygonal karst pattern that Williams (1972) documented on the exposed surface of Miocene limestones in Papua, New Guinea (Fig. 10). Indeed, Purdy and Bertram (1993) used this figure to explain the "peculiar honeycomb pattern" (p. 40) that they observed in shallow lagoon areas in the Maldives. Purdy and Bertram were convinced that preferential reef colonization on polygonal karst relief formed on limestone surfaces during Pleistocene subaerial exposure was responsible for the "honeycomb shoals" (p. 41) in the 8 Maldives. Purdy (1974a, 1974b) also hypothesized that elevated karst Pleistocene relief was responsible for the location and initiation of many of the Holocene lagoon reefs in Belize. TOPOGRAPHIC RIDGES INTERMITTENT STREAM CHANNEL Figure 10. Plan view of polygonal karst pattern on the surface of Miocene limestones, Papua, New Guinea (Purdy and Bertram, 1993; modified from Williams, 1972). Such honeycomb shoals are common features of shallow lagoon areas and have also been well documented in the Cocos (Keeling) Islands (Searle, 1994). Searle noted that the "central southeastern part of the lagoon is occupied by steep-sided “blue holes,’ some over 15 m deep" (p. 5), and 100 m wide (Fig. 11). These, Searle suggested, are related to "multi-generational dolins" (p. 5), although he emphasized that both differential accretion and erosional relief are responsible for atoll lagoon morphology. Figure 11. Aerial view of southern area of Cocos (Keeling) Islands lagoon. Note honeycomb ridge pattern ( Searle, 1994). A core hole drilled in the Pelican Cays by the University of Miami in 1984 recovered Pleistocene limestone at a depth of about 20 m (R. N. Ginsburg, personal communication, 1994). This depth coincides with the depth at which our probe encountered hard rock on the outside of Pond C. It is therefore not unreasonable to assume that our probe was recording the Pleistocene limestone surface substrate when we hit hard rock at the base of our probes. These probes confirm the hypothesis that the Pleistocene karst relief is controlling the honeycomb patterns of reef-ridge growth in the Pelican Cays, which commonly results in the formation of ponded areas. The short cores suggest that Acropora cervicornis colonized the areas of slightly elevated relief on the Pleistocene limestone surface when it was flooded by the rising seas of the Holocene Transgression. Differential growth of this fast-growing branching coral community, reported to be accumulating in this area at a rate of up to 8 m/1,000 yrs by Westphall (1986), then formed steep-sided ridges, which on catching up with sea level in some areas, were overgrown by mangrove communities. Although the ponded network pattern of reef ridges is related to polygonal karst relief on a Pleistocene limestone substrate, their relief is mainly the result of differential reef accumulation (Fig. 5). The origin and growth of these Holocene lagoon reefs is therefore related to a combination of karst control (Purdy, 1974a; 1974b) and differential reef growth (Halley et al., 1977). This honeycomb pattern is superimposed on a larger rhombohedral configuration of the shelf atolls, including the one on which the Pelican cays are located (Fig. 1). The parallelism of these rhomboidal atolls suggests that pre-Holocene reef accumulation occurred along the edges of fault blocks (Purdy, 1974a, 1974b; Precht, 1997). The cores from inside Pond C show a distinct transition from a predominantly Acropora cervicornis facies to a predominantly Porites divaricata facies, dating to probably about 500 yrs B.P. This facies change documents a shallowing-upward reef sequence similar to that reported in this area by Westphall (1986), where a "catch-up" reef community is being replaced by a very shallow "keep-up" community (Neumann and Macintyre, 1985). The Porites divaricata community was killed off on the ridges inside the pond when water conditions became restricted following an almost complete closure of the mouth of the pond. A date of 70 + 60 yrs B.P., if valid, indicates that the present restricted conditions within Pond C are recent. Outside of ponds, Porites divaricata is most common at depths shallower than | m. Radiocarbon dates (Table 2) indicate that red mangrove, Rhizophora mangle, communities became established on the Pelican Cays reef ridges at approximately 600 yrs B.P., and there has been a continuum of mangrove colonization ever since. They can be seen in the process of establishing themselves on some submerged ridges today (Fig. 12). Figure 12. Mangroves at the initial stage of colonization of a shallow submerged reef- ridge crest, Cat Cay. Note the mangrove roots penetrating a surface cover of Porites divaricata, Thalassia testudinum, and Dictyota sp. 10 CONCLUSIONS The Pelican Cays, located in the south-central lagoon of the Belize Barrier Reef, consist of mangrove-covered islands with a network of circular ponds. Probing and coring studies indicate that the ponds are formed by differential sediment accumulation, with the faster accumulating reef facies growing on the positive karst relief on the underlying Pleistocene limestone. Acropora cervicornis-dominated communities, with documented accumulation rates in this area of Belize of up to 8 m/1,000 yrs, have been responsible for accentuating the positive polygonal karst relief. On catching up with sea level, these reef ridges show a transition from an A. cervicornis facies to a very shallow water Porites divaricata facies. Reef growth is terminated when these ridges become shallow enough for colonization by Rhizophora mangle. Islands are eventually formed when these ridges become capped with mangrove peat. The distribution of ponds in the islands reflects the original polygonal relief pattern of the Pleistocene surface. ACKNOWLEDGMENTS Special thanks go to Victoria E. Macintyre Saville for field assistance. We would also like to thank William T. Boykins for his valuable help with graphics and text layout, and Thad Murdock for additional assistance with graphics. Fieldwork for this project was supported by the National Museum of Natural History’s Caribbean Coral Reef Ecosystem Program (CCRE Contribution No. 576). REFERENCES Aronson, R.B., and W. F. Precht 1997. Stasis, biological disturbance, and community structure of a Holocene coral reef. Paleobiology 23:326-346. Halley, R. B., E. A. Shinn, J. H. Hudson, and B. Lidz 1977. Recent and relict topography of Boo Bee patch reef, Belize. Proceedings of the Third International Coral Reef Symposium, Miami, 2:29-35. Neumann, A.C., and I.G. Macintyre 1985. Reef response to sea level rise: Keep-up, Catch-up or Give-up. Proceedings of the Fifth International Coral Reef Congress, Tahiti, 3:105-110. Precht, W. F. 1997. Divergent wrench faulting in the Belize southern lagoon: Implications for Tertiary Caribbean plate movements and Quaternary reef distribution. American Association of Petroleum Geologists Bulletin 81:329-333. Purdy, E. G. 1974a. Reef Configuration: Cause and Effect. Society of Economic Paleontologists and Mineralogists, Special Publication No. 18, 9-76. 1974b. Karst-determined facies patterns in British Honduras: Holocene carbonate sedimentation model. American Association of Petroleum Geologists 58:825-855. Purdy, E. G., and G. T. Bertram 1993. Carbonate Concepts from the Maldives, Indian Ocean. American Association of Petroleum Geologists, Studies in Geology, No. 34, 56p. 1] Searle, D. E. 1994. Late Quaternary morphology of the Cocos (Keeling) Islands. Atoll Research Bulletin No. 401, 13p. Shinn, E. A., R. B. Halley, J. H. Hudson, B. Lidz, and D. M. Robbin 1979. Three-dimensional aspects of Belize patch reefs. American Association of Petroleum Geologists Bulletin 63:528 (Abstract). Westphall, M. J. 1986. Anatomy and history of a ringed-reef complex, Belize, Central America. Master's thesis, University of Miami, Coral Gables, Florida, 135p. Williams, P.W. 1986. Morphometric analysis of polygonal karst in New Guinea. Geological Society of America Bulletin 83:761-796. ots ean \) , \y «6 ATOLL RESEARCH BULLETIN NO. 467 A GENERAL BIOLOGICAL AND GEOLOGICAL SURVEY OF THE RIMS OF PONDS IN THE MAJOR MANGROVE ISLANDS OF THE PELICAN CAYS, BELIZE BY IAN G. MACINTYRE, IVAN GOODBODY, KLAUS RUTZLER, DIANE S. LITTLER, AND MARK M. LITTLER ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 S Northeast Cay Mexico § a) ee 7 . j Gy bé 4 Caribbean i Sea Bird'Cays"4 ys Belize City oe —GGas 17°N Co Cat Cay: i Avigennia Cay . .:- Dangriga 1 oe Guatemala Little Cat Ca s SS Honduras CS. “s, 88 W : ae AEQFHS Lagoon Cays : ay NG Island Figure 1. Index map of the Pelican Cays showing the location of the ponds. A GENERAL BIOLOGICAL AND GEOLOGICAL SURVEY OF THE RIMS OF PONDS IN THE MAJOR MANGROVE ISLANDS OF THE PELICAN CAYS, BELIZE BY IAN G. MACINTYRE,' IVAN GOODBODY,? KLAUS RUTZLER,? DIANE S. LITTLER,’ and MARK M. LITTLER* ABSTRACT A basic description of the geological and biological characteristics of the rims of major ponds in the Pelican Cays is presented. This report is based on a 1994 general survey that was augmented by observations of benthic community populations and systematic collections completed between 1992 and 1997. INTRODUCTION The Pelican Cays (16°39.8'N; 88°11.5'W) are a group of mangrove islands (Fig. 1) forming the northwestern section of the Rhomboid Shoals in the southern lagoon of the Belize Brrier Reef complex (Macintyre and Aronson, 1997). These islands occur on elongate shelf atolls with parallel flanks, whose pattern of formation has been controlled by the underlying tectonics of the region (Purdy, 1974a, 1974b). A striking feature of the cays is their distinct circular ponds, which are the result of differential coral and mangrove growth on a Pleistocene subsurface exhibiting a karst polygonal ring pattern (Macintyre et al., this volume). The Pelican Cays ponds constitute pristine, low-energy, benign, and biologically diverse ecosystems dominated by sessile photosynthetic and filter-feeding populations (Littler and Littler, 1997). Most of the marine species are morphologically delicate and vulnerable to damage from the emissions of boat exhausts; physical disturbances by boat wakes, snorkelers, storms, and sedimentation; and natural (Lapointe et al., 1993) and anthropogenic (Littler et al., 1993) eutrophication. 'Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560-0125. *Department of Life Sciences, University of the West Indies, P.O. Box 12, Kingston 7, Jamaica. ‘Department of Invertebrate Zoology, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560-0163. “Department of Botany, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560-0166. 16 METHODS In June 1994, some members of our group conducted a general survey of the geological and biological characteristics of the rims of the Pelican Cays ponds (Fig. 1). Others in the group collected biological information on these ponds at various periods. Maps were drawn from low- altitude, oblique aerial photographs superimposed on existing nautical-charts. Coordinates were confirmed by a hand-held GPS (Garman GP545, 15- to 100-m accuracy), and depth soundings were taken using a hand-held Scubapro PDS-2 dive sonar and were converted from feet to meters. Between February 1992 and August 1997, samples of pond organisms were systematically collected from mangrove prop roots, seagrass flats, and shallow reefs ridges, as well as from the deeper slopes of seagrass beds and reef ridges. Detailed qualitative observations of populations were made using transect-A methods (Goodbody, this volume; Littler and Littler, 1997, this volume; and Riitzler et al., this volume). Voucher specimens were examined live at the Carrie Bow Cay field station; they were then prepared for microscope study, fixed in 10% formalin-seawater, and preserved in 70% ethyl! alcohol. Sections of the plant materials were made by hand (in the field) or by freezing microtome (in the laboratory). Sections were stained with 1% aniline blue and mounted using a 20% glucose syrup (Karo Syrup, Corn Products, Inc.) solution in distilled water containing a trace of phenol. Sponges were sectioned by hand and the sections were dehydrated, cleared, and mounted in Permount (Fisher Scientific Company); preliminary spicule mounts were made after dissolving representative fragments in 5% sodium hypochlorite. Ascidians were relaxed before fixation by adding menthol crystals to the seawater in observation vessels. The specimens from this study are deposited in the National Museum of Natural History, Smithsonian Institution, Washington, D.C. In this discussion, the ponds in the Pelican Cays are identified by letters, although some of the ponds have names (Littler and Littler, 1997): Pond A = Cat Cay Bay; Pond B = Cassiopea Cove; Pond C = Tony's Lagoon; Ponds E and F = Frenchy's Ponds; Pond G = Great Pond; Pond GG = B’; and Pond J = Little Cat Bay. THE SETTING The Pelican Cays ponds (Fig. 1) have a number of characteristics in common. Each pond has at least one opening to the surrounding seas, and these openings are restricted to various degrees by coral ridges. Each is surrounded in part by mangrove forests composed of the red mangrove Rhizophora mangle and sporadic stands of the black mangrove Avicennia germinans. These forests have formed an underlying peat substrate, which is usually eroded around the perimeter of the ponds and undercut to expose some of the root system of the forest above; the exposed roots may be referred to as "bank roots," in contrast to the adventitious "hanging roots" that grow down from the canopy into the water below. Bank roots, hanging roots, the bank rim base, and the exposed peat bank provide substrates for sessile organisms and thus support flourishing communities of sponges, ascidians, algae, corals, sabellid polychaetes, molluscs, bryozoans, and other organisms. Fine organic sediment, derived from the adjacent mangrove forest floor, lies at the base of most pond rims. Observations of the peat margin at ebb tide revealed that a trickle of suspended organic particles often moves across the peat and settles onto the bases of the ponds. These 17 bottom sediments are so fine that they are easily disturbed by any movement of the water, as indicated by the random movement of a swim fin. The resulting cloud of suspended material eventually settles on and stresses sessile organisms on the mangrove roots or the peat bank. In many locations the base of the rim is composed of Halimeda sediments or accumulations of bivalve shells, depending on the communities flourishing on the adjacent roots and peat bank. In addition, coral debris is exposed at the base of many undercut sections of the peat banks. This debris derives mainly from the reef facies on top of the Pleistocene karst polygonal ridges, and with the mangroves is responsible for the formation of these ponds (Macintyre et al., this volume). Where sediment and light conditions permit, turtle grass, Thalassia testudinum, is a common feature on the bottom around the margins of the ponds. In the center of most ponds, the bottom is covered by an organic-rich mud. THE SURVEY Each survey began at the pond openings and continued along the coastal rim in a clockwise direction. Pond A, Cat Cay (Fig. 2) The western entrance to Pond A is closed by a ridge about 5 m wide across the top; at high tide this ridge crest is only 0.5 to 1.0 m below the surface. The outer slope, which descends to a depth of about 20 m, is covered in large part by a rich growth of Agaricia tenuifolia (Plate Ic),> whereas the crest is dominated by Porites divaricata and abundant filamentous algae, with small amounts of Thalassia testudinum. Down the outer slope, particularly below 3-5 m, the coral cover is interrupted by large reef-like stands of barrel and rope sponges (Xesfospongia muta (Plate 3b), Jotrochota birotulata, and Amphimedon compressa). On the inner side, the ridge slopes down to at least 14 m in the pond. This inner slope too has rich growths of A. tenuifolia, with a few stands of Acropora cervicornis, which in 1994 were partly affected by white-band disease. The Agaricia blades are inhabited by large populations of solitary ascidians (Goodbody, this volume) and sponges, particularly the photosynthetic symbiotic Ulosa funicularis (Plate 3a) (Riitzler et al., this volume). The colonies of A. fenuifolia are delicately balanced on this slope and are easily disturbed. When swimmers just brush by them, the colonies can become dislodged and roll down a talus slope toward the bottom of the pond. Here and there, the colonies have been stabilized by large clusters of Zoanthus sp. and encrusting sponges (Chondrilla cf. nucula) (Plate 3a), which have grown between adjacent coral colonies, bound them together and, in the case of Chondrilla, overgrown live coral surfaces. These closely packed plates of coral colonies also provide a cryptic environment for colonial ascidians, sponges, and other sessile organisms. The depth at the center of this pond is 11 meters. 'Most of this outer slope population experienced bleaching and subsequently died after long exposure to warm stagnant waters that lasted from August to November, 1998 (Aronson et al.in press). eT a) 30 meters its algae ey ascidians corals hi mangrove Vv seagrass sponges Figure 2. Schematic map showin rim of Pond A g the distribution of dominant bottom communities around the 19 Starting from this western entrance and moving around the pond in a clockwise direction, corals (primarily Porites divaricata and Agaricia tenuifolia) and algae dominate the rim, flourishing between and around the mangrove roots for about one-third of the way along the western rim to a point where the mangroves become undercut and expose a coral-rubble base composed mostly of Porites divaricata, Acropora cervicornis, and Millepora sp. With a slight increase in depth, this coral rubble gives way to a muddy Thalassia testudinum bottom. From this point on, bank roots, hanging roots, and exposed peat banks on the western rim have an abundant cover of sponges and ascidians, with some bryozoans and hydrozoans and a few anemones (Lebrunea sp.). The dominant algae are the red Coelothrix irregularis, Acanthophora spicifera, and Spyridia filamentosa; the brown Lobophora variegate, Dictyota spp., and Padina sanctae-crucis; and the green Caulerpa racemosa, C. sertularioides, C. mexicana, C. verticillata, Halimeda opuntia, and Dictyosphaeria cavernosa. Such robust macroalgal populations are indicative of constant, stable, low-nutrient conditions (Lapointe et al., 1993). Unusual species include foliose fleshy red algae (Gracilaria, Meristiella), epiphytic crustose corallines, and the large green sand-dwelling alga Udotea cf. occidentalis. These sessile communities yielded specimens of the gastropod Calliostoma sp. and the flatworm Pseudoceros crozieri, both of which are probably predators on these communities. On this western rim, one of the most abundant colonial ascidians is Cystodytes dellechialei, of which there are at least five color morphs; these colonies live preferentially on the peat bank. Several species of sponges are also common on the western rim, notably Scopalina ruetzleri (Plate 3e), Desmopsamma anchorata (Plate 3c), Mycale laevis (Plate 3c), Jotrochota birotulata, Haliclona manglaris, and Mycale laxissima. lotrochota forms dangling ropes of blackish-green sponge hanging from mangrove roots on the lip of the bank. About two-thirds of the way along this western rim there is a distinct mound of dead Acropora cervicornis located a little distance into the pond. The northern shore is composed of a steep peat bank about | m high, with occasional areas of undercutting. Large populations of the anemones Bartholomea annulata and Aiptasia tagetes occupy the roots and peat bank along this shoreline. Small numbers of the bivalve Chama macerophylla are also found on the roots but not in the quantities seen in Pond C. A luxuriant growth of the alga Lobophora variegata covers this peat bank, and oysters (Crassostrea rhizophorae and Isognomon alatus) are attached to the roots. Both sponges and solitary ascidians are sparse along this north bank, but some species of colonial ascidians, notably Clavelina picta (Plate 2c) and Ecteinascidia turbinata (Plate 2e) are present on hanging roots. At the center of the northern shoreline the pond is close to the open barrier-reef lagoon on the windward side of the island. The mangroves here are thin, and water flows into the pond across this section of mangrove when wind conditions are strong (the prevailing winds are from the northeast). Simple float tests suggest that this inflow is carried into the center of the western rim of this pond, where some of the richest and most interesting populations of ascidians occur. A small recess, or channel, 42 m long occurs in the northeast corner of the pond and leads into it from among the mangroves. The bottom sediments in the recess are sulfurous, which suggests poor circulation. Rich populations of algae, ascidians, and sponges are again abundant on roots and the peat bank along the eastern rim of Pond A, but these filter-feeding communities are not as abundant as on the westward rim. The southern tip of this eastern rim is occupied by a large colony of Montastraea annularis. 20 Lying further southeast is a deep channel, and next to it is a thriving coral community surrounding a small mangrove island. Most of this community consists of Agaricia tenuifolia, Porites divaricata, and Millepora sp., with thick growths of Halimeda between these colonies. Scattered colonies of Acropora cervicornis are also present around this island, along with abundant algae and Thalassia testudinum growing between the shallow root system. Notable algae include three markedly different forms of Codium spp. along the southwest margin and large populations of the red alga Hydropuntia cornea on the shallow flats. An Agaricia tenuifolia ridge connects this island westward to a larger mangrove island. Another rich ascidian community flourishes on the mangrove roots of this island. This Agaricia-dominated community continues westward, connecting with the shallow ridge across the pond mouth. Pond B, Manatee Cay (Fig. 3) The entrance to this small pond faces northeast into the direction of the prevailing winds. It is a shallow, muddy pond with a few corals and octocorals at the entrance, which was partly blocked by a shallow flat with two small islands of Rhizophora mangle. The outer, northern margin of these islets has several large populations of the agar-producing red algae Gracilaria cervicornis, G. mammillaris, and Hydropuntia cornea. The rim of this pond has a curtain of hanging mangrove roots covered by algae and ascidians, including Botrylloides nigrum, Polyclinum constellatum, and Perophora regina; the latter two species have not been recorded in quantity from any other ponds in the Pelican Cays (see Goodbody, this volume). The peat bank around the rim of this pond shows little undercutting, and the bottom is covered with fine sediment, which supports abundant populations of several morphological forms of the large jellyfish Cassiopea xamachana. The brown crown conch Melongena melongena is also common. Ulvalean algae, which indicate eutrophication, are present along the eastern rim. In the southern section, particularly just inside of the entrance, Thalassia testudinum flourishes on the mud bottom, giving way to Syringodium filiforme in the northern part of the pond. Both of these seagrasses have exceptionally long blades (> 1 m). Pond BB, Manatee Cay A small, almost circular pond (about 100 m in diameter) is located southwest of Manatee Cay at 16°39.92'N; 88°11.54'W. A narrow entrance measuring about 20 m across occurs on the western side of the pond. Cutting across the entrance is a ridge with well-developed relief. The crest of this ridge is formed of coral rubble with an abundance of Thalassia testudinum and some Zoanthus sp.; various large anemones are also present, including Stoichactis helianthus and a bright green Actinostella flosculifera partly buried in the sediments. Octocorals, most notably Briareum sp., are scattered across the ridge along with a few solitary ascidians; Polycarpa aurita and Ascidia sydneinsis were recorded during a visit in April 1996. The surrounding sediments contain numerous large burrows, suggesting the presence of stomatopods or calianassids. Just south of the entrance is a spectacular population of the giant paddle-like green alga Avrainvillea asarifolia (with blades measuring 30 cm by 20 cm). Inside the lagoon, the entire rim is surrounded by a curtain of hanging roots. The pond floor here is about 1.5 m deep and slopes down to the center of the pond at an angle of about 45°. Much of the peat bank is eroded, exposing a vertical surface about 1 m high. The bottom of this 32 + o§ @2 P oO & algae ascidians corals octocorals mangrove seagrass sponges Figure 3. Schematic map showing the distribution of dominant bottom communities around the rims of Pond B and Pond D. 22 pond consists of fine sediments that are easily resuspended, and the water is generally murky because of suspended particles. Large schools of small harengulid fish inhabit the area, each attended by a barracuda, and apparently play a role in keeping sediments in suspension. No such schools were observed in the other ponds, except for Pond I and occasionally Pond A. Numerous grey (Lutjanus griseus) and schoolmaster (Lutjanus adopus) snappers can be seen among the roots around the rim of Pond BB. Sessile communities on the roots here are dominated by sabellids, with sponges and ascidians also being common. Notable among the ascidians are colonies of Ecteinascidia turbinata (Plate 2e), also found only in Ponds A and C, and Didemnum psammathodes, also observed only in Ponds B and GG. Some roots are populated by clusters of oysters, Jsognomon alatus and Crassostrea rhizophorae, but no barnacles are present. The feathery green alga Bryopsis plumose is particularly robust (at 15 cm in length), while the hanging roots are dominated by an interesting form of the giant-celled green alga Caulerpa nummularia (a new record for the Caribbean). Pond C, Manatee Cay (Fig. 4) Pond C has by far the most spectacular array of plants and multilayered community of filter-feeding animals in the Pelican Cays, including a great variety of algae, colorful tunicates, sponges, tube worms, bryozoans, bivalves, anemones, corals, and brittle stars. The inside of this pond is teeming with juvenile fish and large tarpon (Megalops atlanticus). A ridge stretches across the mouth of this pond. For the most part, the crest is less than 0.5 m deep and about 5 m wide and is covered by corals—Porites divaricata, Millepora alcicornis, and some Agaricia tenuifolia—interspersed with the seagrass Thalassia testudinum. The outer flank consists of a rich cover of A. fenuifolia (Plate 1c), with lesser amounts of P. divaricata, M. alcicornis, and Acropora cervicornis. In marked contrast, the inner flank (Plate 1d) is dominated by macroalgae, mostly Halimeda opuntia along with some Caulerpa and Dictyota. Sponges such as Anthosigmella varians and Lissodendffyx colombiensis, with a rich variety of brittle stars, are abundant. Some scattered colonies of corals occur along with octocorals and zoanthids. The north and south ends of this ridge are inhabited by large populations of the red agariphyte Gracilaria mammillaris, several other commercially valuable Gracilaria spp. are also present. From the north end of the pond mouth and along the interior rim, the bottom of the pond is covered mainly by Halimeda, Thalassia, and Caulerpa. A solitary coral, Scolymia lacera, was found at a depth of 1 m attached to a fragment of P. divaricata; this species normally lives in depths of 14 to 40 m in Belize (Cairns, 1982). The Halimeda-Thalassia bottom has a rich cover of macroalgae where a coral-rubble ridge meets the western rim. Beyond this area, the mangroves are undercut and the exposed roots support a high diversity of solitary and colonial ascidians and sponges, with many brittle stars and platyctenid ctenophores. Sponges cover the hanging roots as well as bank roots and peat surfaces deep into the undercuts, where there is not enough light for algal competitors. Common massive species include Spongia tubulifera (Plate 3d), Scopalina ruetzleri (Plate 3e), Mycale laevis (Plate 3c), lotrochota birotulata, and species of Ircinia and Xestospongia; encrusting forms appearing in abundance are Mycale microssigmatosa, Clathria venosa, Spirastrella mollis, and Terpios manglaris. Examination of this area of the rim at night revealed an abundance of tentacular structures hanging down into the water; these any : cA oe i algae sabes GR feted ascidians G ig v : corals 26 ee ‘G octocorals Wb 9 a mangrove seagrass | ont ee wb sponges Figure 4. Schematic map showing the distribution of dominant bottom communities around the rim of Pond C. 24 include the feeding tentacles of platyctenid ctenophores and terebellid worms, as well as the arms of various brittle stars. The base of the undercuts consists of coral and molluscan debris in varying amounts of a matrix of mud; this debris provides a hard substrate for abundant growths of solitary ascidians, especially Ascidia nigra and Polycarpa spongiabilis. Much of the molluscan debris comes from bivalves, chiefly Chama macerophylla, which is abundant on the mangrove root system above. Toward the northern end of this pond, the bottom becomes muddier, with molluscan debris scattered here and there and an occasional bottom-dwelling jellyfish (Cassiopea xamachana). In the northwest corner of the pond the peat bank is undercut by about 1.5 to 2.0 m, leaving a lip 0.5-m thick hanging over the pond floor 1.5 to 2.0 m below. This deep undercut is about 20 m long and provides a substrate for many species of ascidians, sponges, and algae. Toward the center of this lip there is a luxuriant growth of the alga Caulerpa cupressoides. Curtains of roots hang on either side of the undercut and are heavily encrusted with sponges and ascidians, notably the colonial ascidian Diplosoma glandulosum. At this point on the northern rim of the pond, the mangrove fringe separates the pond from the waters of the open barrier-reef lagoon by a distance of only 16 m, allowing the prevailing trade wind to push water into the pond. A hanging population of extremely large Avrainvillea digitata appears at the intertidal level of the northern margin. Nearby, on the rim base to the east, gigantic Penicillus pyriformis lives among the blades of Thalassia testudinum. The entire northern rim of Pond C is inhabited by a delicate long-lived community that has experienced little human disturbance, most likely because the shallow ridge across the mouth restricts access and the embayment is so large. This northern section of Pond C is 11 m deep at its center. Mangroves along the eastern rim of the pond are not as deeply undercut and Thalassia testudinum covers a muddy bottom, with only scattered ascidians and sponges found on the mangrove roots. Further south, at a point opposite from the opening, two ridges extend across the pond. These ridges consist of Porites divaricata and Acropora cervicornis rubble with a Thalassia and algal (Caulerpa racemosa and C. cupressoides) cover. Some live Porites divaricata was observed at this location on the eastern rim, along with a variety of ascidians, sponges, macroalgae, and Thalassia. Continuing into the southern embayment, the water reaches a depth of 14 m; the mangroves are not deeply undercut and Thalassia covers a muddy bottom. Occasional zooids of the solitary ascidian Polycarpa spongiabilis occur among the Thalassia, but for the most part the mangrove roots are not heavily encrusted with sessile organisms in this murky embayment, which had a great deal of suspended matter in the water column when first visited on April 27, 1993: Further south, a small section of a coral rubble ridge extends into the pond. This ridge consists of Agaricia sp., A. cervicornis, and Porites divaricata debris. Halimeda is abundant on the mangrove roots, among scattered ascidians and gorgonian corals. Extending around the southern rim of the pond, the mangroves are undercut, revealing a peat cover intermixed with coral debris. The mangrove roots support thick growths of Halimeda along with a few colonies of live Agaricia sp. This undercut rim with coral debris at the base and Halimeda between the roots continues right up to the mouth of the pond, with increasing amounts of live coral, including Agaricia sp. and Porites divaricata. In addition, large colonies of the brown encrusting bryozoan Steginoporella magnilabris are present on the mangrove roots. This southern end of the 25 pond has several small embayments in which the mangrove roots support luxuriant growths of sponges and ascidians. The colonial ascidian Ecteinascidia turbinata (Plate 2e) is common here but is not found in the remainder of the pond. The anemones Stoichactis helianthus and Condylactis gigantea are also present here. These southern embayments are shallow and covered with scattered growths of Thalassia testudinum resting in fine sediments, as well as several species of algae, including Penicillus sp. and Udotea sp. Pond D, Manatee Cay (Fig. 3) The mouth of this small pond is crossed by a sill consisting of coral rubble with a surface community dominated by Thalassia testudinum, Manicina aerolata, Porites porites, Millepora sp., Halimeda sp., octocorals, and zoanthids. The first section of the north rim consists of open mangrove roots covered with the encrusting octocoral Briareum sp.(probably B. polyanthes) and algae, with a bottom cover of Halimeda sp., Thalassia testudinum, and scattered Manicina areolata. A large colony of Siderastrea siderea inhabits the roots. On the north rim, the mangroves are undercut, their roots covered with encrusting octocorals up to a small inlet, where sponges, anemones (Aiptasia tagetes), and oysters (Crassotrea rhizophorae) increase in number. At this point hydroids become more abundant and the muddy bottom is covered with Thalassia testudinum and Halimeda. Further to the east, sponges, anemones, and several species of solitary ascidians become more common on the mangrove roots. Colonial ascidians are absent from this pond. The bottom is an organic-rich mud with continuous cover of Thalassia testudinum. The southern rim is formed of undercut mangroves with sponges and algae (most noticeably Halimeda sp. and Caulerpa sp.) on the roots, with an increasing number of octocorals toward the pond opening. The bottom consists of mollusc and coral rubble in a mud matrix with a cover of Thalassia testudinum. The mangrove roots in this pond are sparsely covered by sessile organisms, the octocoral Briareum sp. being the most common colonist. This is surprising in view of the fact that the eastern rim of this pond is very close to the richly diverse western rim of Pond C, where populations of Chama macerophylla along with sponges and ascidians are so abundant. During a visit to this small pond on April 5, 1994, many jellyfish, Cassiopea xamachana, were observed, along with numerous large burrows in the bottom sediments that suggested the presence of large crustaceans, either stomatopods or calianassids. Pond E, Fisherman's Cay (Fig. 5) The entrance to this pond is a narrow gap in the mangroves in the southwest corner that is just wide enough to for a boat to pass through; a manatee swam out through this channel when the site was visited on April 19, 1994. The ridge at the entrance of this pond is almost 4 m deep at its center, which is greater than the depths of ridges found in the other ponds. This ridge is dominated by Agaricia tenuifolia, with some Porites divaricata and Acropora cervicornis, sponges and Halimeda are packed between the colonies. At depths of 0.5 m, the southern crest 1s populated mainly by Agaricia tenuifolia, whereas the shallow northern crest consists of a sandy bottom with Thalassia testudinum and scattered colonies of Manicina areolata, Millepora sp.., Montastraea annularis, Porites astreoides, and zoanthids. Octocorals grow around the mangrove roots, along with a light cover of hydroids. The roots here are undercut, and the exposed peat has algae * ae ascidians corals a ~ of @ Hp eo «& / AG [We octocorals * prose iG iit ib { Pond E mangrove : seagrass AG ae oe : \? sponges / Gg? N re ou % bul A Oe Figure 5. Schematic map showing the distribution of dominant bottom communities around the rims of Pond E and Pond F. Qi a dense cover of the brown alga Lobophora variegata. Inside this pond and almost around the entire northwest rim, the peat bank is eroded; the vertical face has a relief of about 1.5 m and exhibits relatively little undercutting. This bank provides a substrate for many species of ascidians and sponges. Sponges include many large branching clusters of Aplysina fulva and Artemisina melana, massive Scopalina ruetzleri (Plate 3e), Mycale laevis (Plate 3c), and Spongia spp., and encrusting Chondrilla cf. nucula (Plate 3a) and Mycale microsigmatosa. Immediately inside the entrance and extending along the northwestern rim, there is a diverse community of ascidians, especially Clavelina picta (Plate 2c) on hanging roots, with C. puertosecensis (Plate 2d) being more common on the peat bank. Dense growths of green algae are also present, notably Caulerpa verticillata on the bank and C. racemosa on roots, with Halimeda sp. packed between the root system. The latter contributes a considerable amount of carbonate sand to the pond floor, which is inhabited by several species of solitary ascidians, notably Ascidia interrupta, Polycarpa spongiabilis, and Microcosmus exasperatus (Plate 2f). Approaching the north end of this pond, Halimeda sp. between the roots remains thick, but other algae are also present, including Ventricaria ventricosa. The roots here are covered with numerous sponges, the encrusting bryozoans Trematooecia aviculifer and Stylopoma sp., hydrozoans, and the brown alga Lobophora variegata. A narrow pass at this northern end of Pond E forms a channel into a very small pond. This pass has coral-rubble walls (branching Porites sp.) with a Thalassia testudinum-covered bottom, which supports a large colony of Siderastrea siderea. The relatively rare ascidian Halocynthia microspinosa was seen on two occasions on the peat bank in this pond. Sabellids, probably a species of Branchiomma, are also common. Bivalves, on the other hand, are not common, but small numbers of /sognomon alatus, Crassostrea rhizophorae, and Chama macerophylla occur on roots, and a single specimen of Modiolus americanus was found in the peat bank on the eastern rim in April 1994. Also at that time, a species of Atrina was common in bottom sediments. Numerous large burrows in the sediments indicate the presence of either stomatopods or callianassids. The southeast rim has noticeably more hydrozoans and corals than the northwest rim. The mangrove roots opposite the pond opening are encrusted by large populations of ascidians, sponges, and Lobophora sp., while Dictyota sp. and Porites divaricata occur between the roots. The Thalassia/Halimeda muddy bottom contains a considerable amount of branching coral debris. Some colonies of Montastraea annularis appear close the opening into Pond F, but most of them are dead and extensively bored by Cliona sp. sponges. Algal growth becomes very abundant, with Halimeda sp. now found growing on the mangrove roots. A large dead colony of Siderastrea siderea along with some live colonies of Porites divaricata occurs between the roots. Continuing past the entrance into Pond F, the bottom is covered by Porites divaricata debris along with some live material. The mangrove roots are encrusted with Lobophora variegata, sponges, ascidians, and some octocorals, which become very abundant among the roots near the mouth of this pond. The pond is 11 m deep at its center. Pond F, Fisherman's Cay (Fig. 5) A shallow opening on the southeastern rim of Pond E forms the entrance to Pond F. The bottom here consists of branching coral that creates a ridge across the entrance. A dead and extensively bored colony of Montastraea annularis is present in this pass, and sponges, 28 Millepora alcicornis, and Halimeda sp. are attached to the mangrove roots. The northern rim is characterized by large quantities of Halimeda opuntia, intermixed with the ascidian Perophora carpenteria. Sponges (Scopalina ruetzleri (Plate 3e), Aplysina fulva) are also common. Continuing on to the eastern rim, one finds a very rich Thalassia/Halimeda muddy bottom with patches of Halimeda sand. Sponges, ascidians, Halimeda sp., and some bryozoans flourish on the mangrove roots. All of the mangroves in this pond are undercut. The southern end of this pond has ascidians, sponges, and Halimeda sp. on the roots and the exposed peat wall is covered with Halimeda sp. and shingle-like growths of the brown alga Lobophora variegata along with a few bryozoans. The bottom consists of a muddy Halimeda sand with abundant branching coral rubble. The southwestern rim is dominated by large curtains of the microscopic red alga Ceramium sp. on both hanging and bank roots. Other noteworthy algal populations consist of draped masses of Halimeda opuntia suspended from mangrove prop roots and mound-like colonies of Avrainvillea asarifolia, often overgrown by creeping strands of Caulerpa spp. Sponges are particularly abundant on the undercut mangrove roots of the western rim. Oysters, which are common on the roots of Ponds F, are abundant in this southern rim area, along with some Halimeda sp. on the peat banks. Branching coral rubble has accumulated at the base of the undercut and some coral is embedded in the peat. The bottom continues to be a Halimeda sandy mud. The fauna of this pond is similar to that found on the northwestern rim of Pond E. Scopalina ruetzleri (Plate 3e) is the dominant sponge, and solitary species of ascidian are common, notably Pyura lignosa, Polycarpa spongiabilis, and Ascidia interrupta. Clavelina picta (Plate 2c) and C. puertosecensis (Plate 2d) are common all around the rim, along with small colonies of the ascidians Distaplia corolla and Eudistoma olivaceum. Many species of bivalve molluscs inhabit this pond, including, Chama macerophylla, Brachidontes domingensis, Arca imbricate, Codakia orbicularis, and Spondylus circus. During an April 1994 visit, many specimens of the jellyfish Cassiopeia xamachana were observed on the bottom of this pond. Pond G, Fisherman's Cay This is a large horseshoe-shaped bay at the northern end of Fisherman's Cay. It has a wide opening facing north, which is crossed by a shallow coral rubble ridge with an abundance of Thalassia testudinum containing zooids of the ascidians Polycarpa spongiabilis and Herdmania momus and scattered coral heads. Islets of Rhizophora mangle are established along the western edge of this ridge. Noteworthy features are abundant standing crops of commercial agariphytes, such as Gracilaria cervicornis, G. mammillaris, Meristiella gelidium, and M. echinocarpum, along the northwestern (outer) border of the lagoon among isolated colonies of the fire coral Millepora complanata. Ulva rigida blades are prevalent on nearby islet roots directly beneath a seabird roosting site. Moving clockwise around this pond, immediately inside the entrance in the northeast corner is an eroded peat bank about 1.5 to 2.0 m high with no undercutting. The mangrove fringe at this point is narrow, so the waters of the open barrier-reef lagoon have easy access to the pond. Bank roots are exposed on the peat inside this pond, and there is only a thin fringe of hanging mangrove roots, populated mainly by ascidians and some sponges. These ascidian populations do not exhibit the species richness seen in many other ponds, notably Pond 29) A and Pond C. Other common components of the sessile community on roots are the bivalves Chama macerophylla, Crassostrea rhizophorae, and Ivognomon alatus; the anemones Aiptasia tagetes, Bartholomea annulata, and Stoichactis helianthus; and the barnacle Balanus eburneus. The eroded peat bank continues around the rim of this pond, but as it approaches the southeastern rim it loses its vertical relief and eventually slopes down at a shallow angle into the pond and becomes heavily pitted. At this point sponges increase in number while the solitary ascidians decrease. Thalassia testudinum grows on the bottom around almost the entire rim of the pond, and there are occasional patches of Halimeda sp. and other algae. These plants grow on a narrow sediment bottom that slopes steeply toward the center of the pond to a depth of 10 to 12 m. Numerous colonies of the rare ascidian Pycnoclavella belizeana inhabit the southern rim. The spongy green alga Avrainvillea nigricans forms a sparse aggregation on the shallow peat bank at the southern margin, which is the habitat of gigantic specimens of fan-like Udotea cf. occidentalis extending in a 10 m-long by 1.0 m-wide strip along the rim base. Further back among the shallow roots is an extensive patch of A. digitata. Sessile individuals of Cassiopea xamachana are common, and several individuals of the unusual brown crown conch Melongena melongena are also present. The predominant macroalgae on the roots are Acanthophora spicifera and various forms of Caulerpa racemosa. The creeping alga Coelothrix irregularis forms dramatic neon-blue colonies on submerged fallen trees. The benthic community just beneath the mangrove roots rests on a bivalve-Halimeda-hash substrate dominated by rhizophytic plants (Thalassia testudinum and Caulerpa racemosa covered by large mats of Ceramium sp., Caulerpa mexicana, and Caulerpa sertularioides). Dominant on the peat bank among shallow prop roots are the brown alga Padina gymnospora and the filamentous green Caulerpa verticillata. Continued circumnagivation of the pond rim in the northwest corner leads to islets of Rhizophora mangle established at the western end of the ridge crossing the mouth of the pond. Noteworthy features here are abundant standing crops of commercial agariphytes such as Gracilaria cervicornis, G. mammillaris, Meristiella gelidium, and M. echinocarpum, along the northwestern (outer border) of the lagoon among isolated colonies of the fire coral Millepora complanata. Ulva rigida blades are prevalent on a nearby islet on roots directly beneath a seabird roosting site. Ponds GG, Fisherman's Cay This three-pond complex on the western side of Fisherman's Cay, appears to have experienced considerable degradation since earlier observations (1992-93), primarily as a result of sedimentation (stress) and boat damage (physical disturbance) (Littler and Litter, 1997). However, colorful sponges and macroalgae are still prevalent on the mangrove prop roots of both ponds. The western margin is healthy and contains a mangrove root community dominated by gorgonians, anemones, stony corals, and macroalgae such as the shelf-former Lobophora variegata. Adjacent is a seagrass bed with numerous Bryopsidales, red and green colored species of Laurencia, Coelothrix irregularis, and Ceramium sp., grading into an Agaricia tenuifolia ribbon reef containing an individual barrel sponge Xestospongia muta (Plate 3b) that is > 1 m high. The bed drops sharply to a depth of 24 m. 30 Pond H, North Lagoon Cay This is a large pond with a central depth of 15 to 20 m. Its broad entrance lies on the south side, just behind a ridge of coral rubble, octocorals, and Thalassia testudinum. An eroded peat bank 1.5 to 2.0 m high runs around the entire rim and 1s only slightly undercut except for areas of the western rim, where it is undercut as much as 2.0 m. Present along most of the rim is a narrow curtain of hanging mangrove roots, and the base has extensive beds of Thalassia testudinum. One of the most striking features of this pond is the abundance of solitary ascidians on roots and on the bank, notably Ascidia nigra, A. interrupts, and Microcosmus exasperatus (Plate 2f). With the exception of Distaplia corolla, colonial ascidians were not at all common. The brightly colored species of Clavelina, which were so abundant in other ponds, especially Ponds A and C, were not found. Sponges are also sparse, particularly vigorously growing forms such as Scopalina ruetzleri, which were also common in other ponds. On the other hand, this is one of the few ponds inhabited by large populations of the barnacle Balanus eburneus, which are common along some sections of the rim. Bivalves are also appear on roots and on the peat bank, especially Chama macerophylla, Crassostrea rhizophorae, Ostreaftons, and Isognomon alatus. Broken fragments of bivalve shells attest to the presence of an unidentified predator on these communities. The tests of some ascidians such as Ascidia nigra and Pyura lignosa, were found sliced open, probably a result of trunkfish attacks (Lactophrys sp.). As in other lagoons, juvenile parrotfish are abundant grazers along the root and bank rims. Occasional clusters of the anemones Aiptasia tagetes and Bartholomea annulata occur on the upper surface of the peat bank. Ophiurids are also common on the bank and roots around the rim of this pond. The southeast corner has a deep but narrow side channel. The fauna in this channel appears more diverse than in the main pond and includes a number of the colonial ascidians on roots recorded elsewhere in this pond. Sponges are also abundant in this channel, in contrast to the main pond, where they are scarce. The fact that this small side channel is fairly sheltered may account for the abundance of colonial ascidians and sponges. Pond I, South Lagoon Cay This is not so much a pond but a large embayment on the south side of the cay. When visited in April 1993, the bay appeared eutrophic green in color, and it contained large schools of small harengulid fish; the food chain is probably supported by a colony of pelicans nesting in trees in the southeast comer of the bay, which deposits a large supply of fecal material into the water. The roots around this bay have large quantities of the alga Caulerpa sertularioides, an abundance of sabellid worms, many barnacles (Balanus eburneus), and the bryozoan Schizoporella sp. The ascidian fauna is typical of fairly eutrophic conditions and includes Polyclinum constellatum, Symplegma brakenhielmi, Ecteinasciditi styeloides, and Lissoclinum fragile. Ascidia nigra and Ecteinascidia turbinata (Plate 2e), both normally expected to occur in such high-nutrient conditions, are not present. Pond J, Little Cat Cay The entrance to this pond lies on the southern side of Little Cat Cay and sits behind a ail shallow coral rubble ridge containing prominent growths of Thalassia testudinum along with large growth forms of Udotea cf. occidentalis. The pond is small, approximately 175 m long, about 100 m across, and not more than 6 m deep; coordinates for this pond are 160°39.13'N; 88°11.96'W. The pond is entirely surrounded by mangrove forest, and the peat banks are extensively undercut in places, especially along the northeastern rim. This undercutting has exposed a system of bank roots, which are shielded around most of the pond by a fringe of hanging roots; shell and coral debris line the pond floor beneath the undercut. As in other ponds, these roots are covered by extensive growths of sessile organisms, especially the solitary ascidians Pyura lignosa, Polycarpa aurita, Ascidia interrupts, Polycarpa spongiabilis, and Microcosmus exasperatus (Plate 3f), as well as many colonial ascidians, including Botrylloides nigrum, Symplegma brakenhielmi, Lissoclinum fragile, Diplosoma glandulosum, and Perophora carpenteria. Other common inhabitants of the sessile communities are sponges, small anemones, sabellid worms, and the oyster /sognomon alatus. The alga Caulerpa racemosa is common on some roots along the western shore of this pond. Characteristic Communities around Some Cays Cat Cay South Island (see Littler and Littler, 1997), a small mangrove island just below the mouth of Cat Cay Bay, contains three markedly different growth forms of Codium spp. along the southwest margin. Cat Cay South Island is bordered by a narrow grass flat (Thalassia testudinum), except for regular intervals of "ribbon reefs" of the leaf coral Agaricia tenuifolia, which forms crescents connected at their proximal ends to the mangrove stands. Here the trees and seagrass flats extend as points into the ribbon reefs. Several of these extended flats support large populations of the red alga Hydropuntia cornea (which is locally valued as an aphrodisiac when prepared in porridge and is commercially bottled as "Double Trouble Sea Moss" in the Lesser Antilles). These mangrove root systems also contain unusually extensive colonies of stony corals along with abundant macroalgae, sponges, tunicates, and anemones. Avicennia Cay, a small mangrove island resting on an isolated lagoonal patch reef just to the southwest of Manatee Cay, was surveyed from the reef slope upward to the edge of the mangrove prop roots. The steep outer slope was found to be covered by Agaricia tenuifolia leveling off in the shallows to form a gorgonian/seagrass community. A passing manatee was observed in this habitat during May 1995. The north side of the cay is the site of a shallow reef crest formed mainly by the fire coral Millepora complanata and the finger coral Porites porites. Behind this crest, shallow seagrass beds adjoin the mangrove treeline. Agariphytes (Gracilaria cervicornis, G. mammillaris, and Hydropuntia cornea) are abundant on the seagrass and mangrove root systems here. The slight embayment on the south end of the cay is sandy and slopes upward to dense stands of Thalassia testudinum; the mangrove root systems are not particularly rich, and their algal populations are similar to those on the west side of Steward Cay to the north. Ridge Cay is surrounded by seagrass beds and is the only site containing the economically important carrageenan producer Eucheuma isiforme (also used to make a sea-moss beverage). The eastern-facing (outer) bedded mangrove prop roots are covered with interesting and colorful populations of small epiphytic clumps of seaweeds such as the grape-like Botryocladia spinulifera, mesh-like Anadyomene stellata, feathery Bryopsis plumosa, Laurencia spp., Codium spp., and encrusting Peyssonnelia boergesenii. Some lagoonward hanging roots 32 support unusually large populations of various Caulerpa species and calcareous Halimeda opuntia. Northeast Cay, the northernmost island, contains a substantial fish camp at the western tip and serves as a popular anchorage for sailboats. The small coves southeast of the fish camp contain a community similar to the GG ponds of Fisherman's Cay, but populations are small, and the area is overgrown by curtain-like masses of a colonial Ceramium diatom. The mangrove prop roots outside the coves are richer but red-algae are predominant (Laurencia papillose, Digenia simplex, Acanthophora spicifera). Unattended fish traps inside the small coves during May 1995 contained dead and dying French angel fish and gray snapper. The northern margin of this island is especially rich in gorgonian corals (fan and branched forms) and elkhorn coral (Acropora palmata) patch reefs, with another developing fish camp conspicuous along the shoreline. The Bird Cays consist of three islands, the westernmost of which has an algal community similar to that of Co Cat Cay and Little Cat Cay to the south (described below). The northeastern and southeastern Bird Cays are located in the atoll-like lagoon and are surrounded by narrow seagrass flats that drop from an Agaricia margin to steep sediment-strewn slopes. Both of these islands contain bird roosting sites; ulvalean green algae associated with eutrophication are present but minimal. Populations at Co Cat Cay and Little Cat Cay, the two westernmost cays in the Pelican Cays group, are not particularly noteworthy except for their algal floras, which resemble those of the western cays of Blue Ground Range and Stewart Cay to the north. The eastern borders of each are composed of narrow strips of seagrass beds (Thalassia testudinum) that slope steeply into the lagoon with scattered heads of the branched fire coral Millepora complanata and the massive starlet coral Siderastrea siderea at the upper margins. The mangrove-root epiphytes are heavily laden with sediment and consist of various Galaxaura spp., Acanthophora spicifera, and Lobophora variegata. The western borders are richer and more interesting, with broad shallow seagrass flats containing large fan and whip gorgonians, Millepora complanata, Montastraea annularis, and Porites heads. The roots are rich in Anadyomene stellata, Lobophora variegata, Gracilaria mammillaris, Bryopsis plumose, and many Caulerpa spp., along with anemones and corals. The southern portion outside the entrance to Pond J is especially rich in seaweeds such as Codium sp. and Hydropuntia sp. of impressive size. The mangrove prop roots on the western tip of Little Cat Cay are heavily epiphytized by the weedy red alga Acanthophora spicifera. The outer (eastern) margin of the reef at the southeast tip of Cat Cay, just at the start of Five Mile Flat (see Littler and Littler, 1997) where it adjoins the island, was surveyed down to depths of 30 m. Below a depth of 26 m, the plan sandy bottom levels off and gradually slopes to deeper depths. From depths of 20 to 26 m, the bottom consists of sand-inundated rubble with sparse biota. From depths of 20 to 8 m, Halimeda spp. and Lobophora variegata are abundant. From depths of 8 m and upward, sponges, bryozoans, and the green macroalgae Penicillus pyriformis, Halimeda spp., Udotea spp., and the brown L. variegata are predominant among the Agaricia tenuifolia that flourishes between 3 and 15 m. Many calcareous algae such as Peyssonnelia boergesenii, Jania adherence, crustose Corallinaceae, and Amphiroa spp. are present on the dead skeletons of A. tenuifolia. In depths of less than 7 m, the seagrass Thalassia testudinum, large gorgonians (whips and fans), and corals such as Porites porites, P. astreoides, and Millepora complanata are prominent. 33 DISCUSSION The preceding observations are of a general qualitative nature owing to the lack of good quantitative data on most of the communities. Furthermore, many organisms were difficult to identify accurately in the field. Nevertheless, the survey yielded sufficient information demonstrating that the Pelican Cays ponds are of considerable interest and can contribute greatly to scientific understanding of Caribbean marine sessile communities. A challenge for future studies will be to explain the differences and similarities between the various ponds. For example, the western rim of Pond A is similar in many respects to the northwestern part of Pond C, but both differ markedly from the western and northern rims of Pond H. All three of these ponds are large and are oriented in a north-south direction, so they remain sheltered from the prevailing northeast trade winds. All three ponds, however, show some sign of water replacement at the northern end across the mangrove rim when the tide is high and the trade winds are blowing strongly. Whether this water flow pattern influences the types of communities in the ponds is at present unclear. Several of the ponds exhibited high levels of species diversity, notably Ponds A, C, and E; this may be associated with the proximity of coral reef communities at the pond entrance. Some other ponds have unique features: Pond D, for example, is populated by octocorals that do not grow to any extent in the other ponds; Pond B has populations of Perophora regina, an ascidian otherwise recorded only from Twin Cays and Blue Ground Range further north on the barrier reef platform. Pond BB, which is very close to Pond B and separated from it by only a narrow area of mangrove swamp, lacks Perophora regina but has an abundance of another perophorid, Ecteinascidia turbinata, otherwise abundant in only in Pond C. Balanus eburneus, a large barnacle widely distributed in Caribbean mangrove communities, is only sparsely distributed in the ponds of the Pelican Cays. It occurs in Ponds C, G, H, and L, but not in Ponds A or E. Sponge diversity in the better-studied ponds A, C, E, and F is different in species spectrum from all others mangrove species studied in the Belize lagoon and approaches that of open reefs, although the species composition only partly overlaps that of known reef environments. About 40% of sponge species collected in the Pelican Cays ponds are unusual, aberrant, or undescribed. In summary, the Pelican Cays represent a spectacular, high-biodiversity, low-energy environment dominated by photosynthetic and filter-feeding populations. Most are physically delicate and highly susceptible to damage from boat wakes, careless swimmers, physical contact (e.g., trampling), sedimentation, and nutrient enrichment. The existence of such high biodiversity in a small geographic area can be attributed to the unique juxtaposition of mangrove, coral, seagrass, and algal biomes under stable oligotrophic conditions (as indicated by the consistently "gin-clear" waters). Few of the ephemeral sheetlike and filamentous green algae indicative of eutrophic bird islands (e.g., Man-of war Cay, Douglas Cay) or anthropogenically polluted systems (see Littler and Littler, 1997) are present. A long-term study of the trophic structure of the Pelican Cays ponds could shed some light on the factors controlling the distribution of organisms within them. 34 ACKNOWLEDGMENTS Special recognition is due to Paul Shave of Wee Wee Cay who first introduced us to the Pelican Cays by showing us Pond A in 1992. That visit marked the beginning of our studies in this area. Tony Rath brought to our attention Pond C and provided some excellent aerial photographs of the ponds. Very special thanks to Victoria E. Macintyre Saville, Michael R. Carpenter, and Bruno Pernet for field assistance. We are also grateful to William T. Boykins and Molly K. Ryan for valuable help with graphics and design; Barrett L. Brooks for support during botanical work; and Charlotte Goodbody for assistance in the preparation of this manuscript. Funding for the botanical research was provided by NSF Grant DEB-9400534, Harbor Branch Oceanographic Institution, and the Smithsonian Marine Station, Fort Pierce, Florida. Fieldwork for this survey was supported by the National Museum of Natural History's Caribbean Coral Reef Ecosystem Program (CCRE Contribution No. 577). REFERENCES Aronson, R. B., W. F. Precht, I. G. Macintyre, and T. J. T. Murdoch 2000 Coral Bleach-Out in Belize. Nature (in press). Cairns, S. D. 1982. Stony corals (Cnidaria: Hydrozoa, Scleractinia) of Carrie bow Cay, Belize. In The Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize, edited by K. Riitzler and I. G. Macintyre, 271-302. Smithsonian Contributions to the Marine Sciences, No. 12. Washington, D.C.: Smithsonian Institution Press. Lapointe, B. C., M. M. Littler, and D. S. Littler 1993. Modification of benthic community structure by natural eutrophication: The Belize Barrier Reef. Proceedings of the 7th International Coral Reef Symposium, Guam, v. 1, pp. 323-334. Littler, D. S., and M. M. Littler 1997. An illustrated marine flora of the Pelican Cays, Belize. Bulletin of the Biological Society of Washington 9:1-149. Littler, M. M., D. S. Littler, and B. E. Lapointe 1993. Modification of tropical reef community structure due to cultural eutrophication: The southwest coast of Martinique. Proceedings of the 7th International Coral Reef Symposium, Guam, v. 1, pp. 335-343. Macintyre, I. G., and R. B. Aronson 1997. Field guide to the reefs of Belize. Proceedings of the Sth International Coral Reef Symposium, Panama, v. 1, pp. 203-222. Purdy, EsG: 1974a. Reef configuration: Cause and effect. Society of Economic paleontologists and Mineralogists, Special Publication, No. 18, pp. 9-76. 1974b. Karst-determined facies patterns in British Honduras: Holocene carbonate sedimentation model. American Association of Petroleum Geologists 58:825-855. 35 APPENDIX STATUS OF SCLERACTINIAN CORALS IN SOME PELICAN CAYS PONDS FOLLOWING THE 1998 BLEACHING EVENT BY IAN G. MACINTYRE On June 11, 13, and 15, 1999, I visited Pelican Cays ponds A, C, D, and E to assess the status of the scleractinian corals following the massive 1998 bleaching and die-off of corals in the surrounding rhomboid shoals (Aronson et al., 2000). Unusually high water temperatures recorded on the shallow reefs during this bleaching killed almost all the agariciid corals in the shoals and the few Acropora cervicornis that had survived an earlier outbreak of white-band disease (Aronson and Precht, 1998). Pond A, Cat Cay All the Agaricia tenuifolia colonies that previously formed a rich cover on both flanks of the ridge across the western entrance of this pond are now dead, as are the associated Acropora cervicornis, Millepora alcicornis, and most of the Porites divaricata. Much of this recently dead coral is already extensively covered by the encrusting sponge Chondrilla cf. nucula, mats of Zoanthus sp., and macroscopic algae, primarily Halimeda opuntia (Figs. 1, 2). Only traces of live Porites divaricata, Colpophyllia natans, and Porites astreoides were found on the crest of this ridge (Fig. 3a,b). No live corals were found along the western rim north of this entrance. The once lush growths of Agaricia tenuifolia and Porites divaricata around the red mangrove stilt roots are dead and overgrown by sponges, zoanthids, and macro-algae (Fig. 4). The other coral area in Pond A occurs at the southern end, where coral communities once formed a rich cover around the small mangrove island and along the ridge connecting this island to the western entrance ridge. All Agaricia tenuifolia, Porites divaricata, Millepora sp., and Acropora cervicornis reported in our original survey are dead and covered with an expansive growth of sponges and macro-algae such as Halimeda opuntia and Caulerpa racemosa and clumps of blue-green algal filaments (Fig. 5). Pond C, Manatee Cay Here, too, the spectacular growths of Agaricia tenuifolia with some Porites divaricata, Millepora alcicornis, and Acropora cervicornis that covered the outer slope of the ridge across the western entrance are dead and partly covered by the encrusting sponge Chondrilla cf. nucula (Fig. 6). The inner slope has an even denser cover of C. cf. nucula, Zoanthus sp., and Halimeda opuntia over coral rubble. The crest supports a few live colonies of P. divaricata being encroached upon by zoanthid mats and encrusting sponges (Fig. 7). A quick check for the live P. divaricata reported on the eastern rim opposite the entrance 36 revealed only dead colonies, still in growth position and encrusted by C. cf. nucula. However, a few small live colonies of Porites astreoides were found attached to mangrove roots. On the southern rim, only traces of live P. divaricata and P. astreoides were found, now overgrown by vigorous colonies of Zoanthus sp. (Fig. 8). Pond D, Manatee Cay The Thalassia-covered low ridge across the western entrance to this pond still supports healthy colonies of Manicina aerolata (Fig. 9) and Porites divaricata. Along the north rim, P. divaricata is struggling to survive an overgrowth of zoanthids, Halimeda opuntia, and anemones (Fig. 10). The large Siderastraea siderea colony earlier reported on this north rim is still alive but clearly competing with zoanthid mats, encrusting C. cf. nucula, and algal growth (Fig. 11). High turbidity inside this pond impeded observations. Pond E, Fisherman's Cay The shallow north end of the ridge across the entrance to this pond still supports live colonies of Porites divaricata (Fig. 12a) and Manicina aerolata on a sandy Thalassia-covered bottom. By contrast, the deeper southern crest, which was originally covered by a flourishing coral community dominated by Agaricia tenuifolia, has no live corals. Most of the dead colonies are still in growth position and are encrusted by the sponge Chondrilla cf. nucula and overgrown by Halimeda opuntia (Fig. 12b). Small live colonies of P. divaricata on peat banks and P. astreoides on mangrove roots were found on both the southern and eastern rims. The large colony of Siderastraea siderea previously reported from the north end of the pond is still flourishing there. DISCUSSION Though brief, the visits to Ponds A, C, D, and E in 1999 revealed that the 1998 bleaching event caused major changes in the benthic communities of the Pelican Cays ponds. There has been a drastic loss of coral cover, with only a few scattered live colonies of hardier species surviving, including Porites divaricata, P.astreoides, Colpophyllia natans, Siderastraea siderea, and Montastraea annularis. The most vulnerable corals have been Agaricia tenuifolia, Acropora cervicornis, and the hydrocoral Millepora alcicornis: no live colonies of these species were found in the ponds or on the outer flanks of ridges across the entrances. Sponges, colonial zoanthids, and algae have benefited from this demise of the corals and are flourishing on the recently dead skeletons. Some are even growing over the coral remnants that are still alive. Among the sponges, the encrusting Chondrilla cf. nucula is the most noticeably abundant organism, accompanied by extensive mats of Zoanthus sp. and heavy growths of the green algae Halimeda opuntia and Caulerpa racemosa. More detailed observations planned for the near future will document this change more accurately and possibly provide a sense of the long-term trends in the species composition of the benthic communities of the Pelican Cays ponds. Sf REFERENCES Aronson, R. B., and W. F. Precht 1997. Stasis, biological disturbance, and community structure of a Holocene coral reef. Paleobiology 23:326-346. Aronson, R. B., W. F. Precht, I. G. Macintyre, and T. J. T. Murdoch 2000. Coral Bleach-Out in Belize. Nature. (In press). 38 Figure 1. Pond A - Cat Cay. Outer flank of ridge across western entrance. Dead Agaricia tenuifolia extensively encrusted by Chondrilla cf. nucula and surrounded by growths of Halimeda opuntia. Scale = 20 cm. June 13, 1999. Figure 2. Pond A - Cat Cay. Inner flank of ridge across western entrance. Very extensive cover of encrusting sponge Chondrilla cf. nucula over coral rubble. Note scattered growths of Halimedia opuntia and abundance of brittle stars (arrows). Scale = 20 cm. June 13, 1999. Figure 3. Pond A - Cat Cay. Crest of ridge across western entrance. A) Patches of live Porites divaricata on a bottom consisting of dead coral encrusted by Chondrilla cf. nucula. B) Live colony of Colpophyllia natans surrounded by dead coral encrusted by C. ef. nucula and packed with Halimeda opuntia. Scale = 20 cm. June 13, 1999. 39 40 Figure 4. Pond A - Cat Cay. Western rim north of the western entrance. Dead Agaricia tenuifolia partly covered by Chondrilla cf. nucula, Halimeda opuntia, and mats of colonial Zoanthus sp. Scale = 20 cm. June 13, 1999. Figure 5. Pond A - Cat Cay. Formerly coral-rich area around a small mangrove island at the southern end of this pond. Dead Agaricia tenuifolia being overgrown by Halimeda opuntia and tufts of filamentous blue-green algae. Scale = 20 cm. June 13, 1999. 41 Figure 6. Pond C - Manatee Cay. Outer flank of ridge across western entrance. Dead coral bottom -- mostly Agaricia tenuifolia with some Acropora cervicornis and Millepora alcicornis. Patchy encrustations of Chondrilla cf. nucula (arrows). Scale = 20 cm. June 13, 1999. Figure 7. Pond C - Manatee Cay. Crest of ridge across western entrance. Live Porites divaricata almost smothered by encrusting colonial Zoanthus sp. and Chondrilla cf. nucula. June 15, 1999. 42 Figure 8. Pond C - Manatee Cay. Southern rim of this pond. Live Porites astreoides almost overgrown by colonial Zoanthus sp. June 15, 1999. Figure 9. Pond D - Manatee Cay. Crest at north end of ridge across entrance. Live Manicina aerolata (arrows) on a Thalassia/coral rubble bottom. June 15, 1999. 43 ete sae BRR RAN Re Figure 10. Pond D - Manatee Cay. North rim of this pond near the entrance. Tips of live Porites divaricata poke up through a cover of colonial Zoanthus sp., Halimeda opuntia, and an anemone. June 15, 1999. Siderastraea siderea coping with attached Zoanthus sp., Chondrilla cf. nucula (arrow), and macroscopic algae. June 15, 1999. 44 Figure 12. Pond E - Fisherman's Cay. Crest of ridge across entrance. A) Shallow north end of ridge. Live Porites divaricata on a Thalassia/Halimeda bottom. B) Deeper south end of this ridge. Dead Agaricia tenuifolia, mostly in growth position, encrusted by Chondrilla cf. nucula and surrounded by Halimeda opuntia. June 1S, 1999: ATOLL RESEARCH BULLETIN NO. 468 MANGROVE PEAT ANALYSIS AND RECONSTRUCTION OF VEGETATION HISTORY AT THE PELICAN CAYS, BELIZE BY KAREN L. MCKEE AND PATRICIA L. FAULKNER ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 —_—__ km Carbonate Shoals (<2m depth) Mexico Caribbean Sea Northeast Cay ; EC. Honduras (e Aa 88°W X sy Manatee Cay We) pe & Cat Cay _ 2 Guatemala y awe 7m PELICAN ® CAYS Figure 1. Index map showing the location of the Pelican Cays in the Belizean Barrier Reef Complex. Modified from a Landsat TM image acquired 18 September 1987. MANGROVE PEAT ANALYSIS AND RECONSTRUCTION OF VEGETATION HISTORY AT THE PELICAN CAYS, BELIZE BY KAREN L. McKEE! and PATRICIA L. FAULKNER? ABSTRACT The substrate beneath mangrove forests in the Pelican Cays complex is predominately peat composed mainly of mangrove roots. Leaves and wood account for less than 20% of the peat mass. At Cat Cay, the depth of the peat ranges from 0.2 m along the shoreline to 1.65 m in the island center, indicating that the island has expanded horizontally as well as vertically through below-ground, biogenic processes. Mangrove roots thus play a critical role in the soil formation, vertical accretion, and stability of these mangrove cays. The species composition of fossil roots changes markedly with depth: Rhizophora mangle (red mangrove) was the initial colonizer on a coral base, followed by Avicennia germinans (black mangrove), which increased in abundance and expanded radially from the center of the island. The center of the Avicennia stand ultimately died, leaving an unvegetated, shallow pond. The peat thus retains a record of mangrove development, succession, and deterioration in response to sea-level change and concomitant hydroedaphic conditions controlling dispersal, establishment, growth, and mortality of mangroves on oceanic islands in Belize. INTRODUCTION Because the mangrove-dominated islands in the Pelican Cays group (Fig. 1) occur many kilometers from the Belizean mainland and from sources of terrigenous sediments, vertical accretion has occurred primarily through autochthonous processes. Mangrove-dominated areas at the Pelican Cays and elsewhere in the Belize Barrier Reef system are underlain by deposits of peat, in some cases several meters in depth (Cameron and Palmer, 1995; Macintyre et al., 1995; Macintyre et al., this volume). These conditions suggest that a major process contributing to vertical accretion has been organic matter deposition, as reported for Grand Cayman (Woodroffe, 1981), Bermuda (Ellison, 1993), and other mangrove forests located within carbonate settings (Parkinson et al., 1994). Therefore, one objective of this study was to investigate the composition of peat deposits in the Pelican Cays complex and to quantify the relative contribution of mangrove organic matter to the vertical growth of these islands. A second objective was to identify fossil plant fragments in mangrove peat and then reconstruct the vegetation history at the Pelican Cays. Current vegetation patterns and physico-chemical conditions were also examined 'USGS-National Wetlands Research Center, 700 Cajundome Blvd., Lafayette, LA 70506. “Wetland Biogeochemistry Institute, Louisiana State University, Baton Rouge, LA 70803. 48 to support interpretation of vegetation change in these island forests. METHODS This study was conducted at Cat Cay, in the Pelican Cays group (Fig. 1). This island currently has a well-developed mangrove community comprised of three species: Rhizophora mangle (red mangrove), Avicennia germinans (L.) Stearn (black mangrove), and Laguncularia racemosa (L.) Gaertn. f. (white mangrove). A 270-m transect was established across Cat Cay (Fig. 2), in an east-west orientation, and forest composition and soil characteristics were measured at 10-m intervals. At each sampling station, maximum heights of the mangrove species present in the canopy were measured with an extendible rod, and their relative abundance was estimated visually. Soil variables (redox potential [Eh] at 1- and 15-cm depths, porewater sulfide, salinity, pH, and concentrations of NH,- N and PO,-P) were determined as described previously (McKee et al., 1988; McKee, 1995). Soil cores (50 cm’) were also collected for the determination of bulk density, percent ash, and percent organic matter. @ Core sites Mangrove EF Dieback Figure 2. Map showing transect and core locations across Cat Cay. 49 To assess past vegetation composition in relation to current zonation patterns, deeper peat cores were collected along a portion of the transect, traversing vegetation zones (Fig. 2). Cores were taken with a Russian peat sampler, which caused no vertical compaction, to a maximum depth of 2 m and were divided into 10-cm sections (Fig. 3). Each peat section was initially examined for degree of decomposibility (von Post Pressing technique; Parent and Caron, 1993), color, structure (e.g., crumbly, fibrous, etc.), and presence of live roots. A subsample (ca 1 g fr wt) was removed from each peat section and gently washed with seawater through a 2.0-mm and a 0.1-mm mesh sieve. All fragments retained on the 2.0-mm mesh sieve were examined under magnification and separated by type (root, leaf, wood, or other origin). Root fragments were identified to species using a key based on root microscopic and macroscopic features (McKee, unpublished data). The identified components were dried at 65°C and weighed. Fragments retained on the 0.1-mm mesh sieve were also quantified and examined for composition. Unsieved subsamples were analyzed for percent ash and organic matter using the loss-on-ignition method. Figure 3. A 0.5 m section of a mangrove peat core and the Russian peat sampler. RESULTS The mangrove vegetation at Cat Cay was characterized by spatial zonation patterns 50 generated by variation in relative abundance of the two dominant species. The island periphery was dominated by a narrow fringe of R. mangle, followed by a landward zone of A. germinans (mixed or monospecific) and an interior pond devoid of emergent vegetation (Fig. 4). The pond was shallow (< 0.3 m) and contained remnant stumps and roots of A. germinans (Fig. 5). Laguncularia racemosa (up to 15 m high) occurred as scattered individuals about 20 m from the shoreline. > 100 oS — Q 5 s q z Rhizophora 5 50 Hi Avicennia < & [L) Laguncularia 3 25 ra * Unvegetated fa 0 (>) >) =) =) N + oO N 250 270 KR S Distance (m) Figure 4. Relative abundance of mangrove species in the forest canopy across Cat Cay. Figure 5. Remnant stump (left) and cable roots (above) of Avicennia trees in a dieback area. 51] 10 Tree Height (m) Fringe Scrub Woodland Unvegetated Woodland Fringe Salinity (%o) Eh (mV) Depth (cm): =o jl =@— 15 Sulfide (mM) 100 —O— NH4-N (uM) —®— PO4-P (uM) Nutrients (Porewater) 1) oOo Distance (m) Figure 6. Variation in mangrove tree height and soil physico-chemical factors across Cat Cay. 52 Peat component: 0.0-0.1 ALLE Depth (m) 50 I) Percent of Total Mass Figure 7. Variation in peat composition with depth. Relative proportions by mass of organic and inorganic components (average of seven cores) (A). Relative E] Inorganic Ei Organic 100 proportions by mass of fragments separated by size (B) and type (C) from a representative core collected in the dieback area. Nonroot fragments were of leaf or wood ongin. Depth (m) Fragment E) Coarse B. size: 0.0-0.1 0.1-0.2 0.2-0.3 0.3-0.4 0.4-0.5 0.5-0.6 0.6-0.7 0.7-0.8 0.8-0.9 0.9-1.0 & Fine 0 ES) 50 1 100 Percent of Total Mass Fragment ] Root C; ae — Nonroot 0.0-0.1 0.1-0.2 4: 0).2-0.3 0.3-0.4 Fins: 0.4-0.5 0.5-0.6 Fis 0.6-0.7 4: 0.7-0.8 0.8-0.9 4:: 019-10 sis: 0 25 50 75 100 Percent of Total Mass 33 ‘sisAyeur yeod uo paseq AeZ yeD Je AlO\sTY UONL}OSOA Jo UONONSUOIOY *g oins1y ee Ee A1O\SIF{ UOTRIAB9A_ JO uononnsuosaY DSOWAIDA ajsupu SUDUIULAS piuvjnounsv7T ~— vaoydoziyy DIUUAIIAY us ep Uo j@10o BS pues — 00O1?0 Ea SL:°St faa OS:0S ge St:SL x 0:00. zz sully puv[poo Pp Lee ee ee eee 00 (wi) yidaq Id ayejadsaauy, 091 Aes kD 54 Soil physico-chemical conditions varied along the transect and in conjunction with changes in vegetative cover (Fig. 6). The soil (0- to 15-cm depth) at all sampling stations was peat (organic matter content 50-60% by mass). Algal-bacterial mats formed 0.25-m plates on the soil surface of the pond. Salinity varied from 32%o at the shoreline to 62%o at the east side of the pond. Rhizophora-dominated zones were generally characterized by lower salinity, whereas Avicennia-dominated zones and the unvegetated area were hypersaline. The vegetated soils were generally more oxidized with low sulfide, compared to the unvegetated pond where sulfide concentrations reached 2 mM (Fig. 6). Concentrations of readily available nutrients also varied across the transect, but did not differ between vegetated and unvegetated areas (Fig. 6). The peat cores collected with the Russian peat sampler were predominately organic, except the basal sections where deposits of calcareous sand and coral were encountered (Fig. 7). On a mass basis, organic matter accounted for 50 to 60% of the cores, indicating a peat-based soil throughout. Except for the surface 30 cm where the majority of live roots occurred, the peat was moderately decomposed (#5 on the Von Post scale), somewhat pasty, and with distinct plant residues. Color ranged from reddish-brown to grayish-brown with blackened particles, and texture varied from fibrous to crumbly. Coarse fragments (> 2 mm) accounted for 25 to 50% of the peat cores (Fig. 7). Most (50 to 95%) of this coarse fraction consisted of mangrove root epidermis or small root sections (Fig. 7), whereas the fine fraction was mostly fine roots (0.1 to 0.2 mm in diameter) and fragments of spongy cortex from larger roots. The surface 20 cm of dieback cores contained more leaf and wood fragments than did deeper layers (Fig. 7). However, leaf and wood fragments decreased in abundance with depth and were minor components in peat overall (Fig. 7). The root fragments were readily identified to species with the key and revealed that the bulk of the peat was composed of Rhizophora and Avicennia roots. Stratigraphy of cores collected in the dieback area at Cat Cay revealed a distinct change in dominant vegetation over time (Fig. 8). Basal sections of peat consisted solely of Rhizophora root fragments and fine roots and occasional leaf fragments. At a 1-m depth, Avicennia root fragments were encountered, and these increased in abundance to the surface layers, which were dominated by this species. DISCUSSION The vegetation and soil conditions currently found on Cat Cay are similar to those of other mangrove islands in the Belize Barrier Reef complex (McKee, 1995; Woodroffe, 1995). Variation in species composition and occurrence of dieback areas in association with changes in hydroedaphic conditions suggest that the zonation of these forests is strongly tied to processes controlling surface elevations in relation to mean sea level (MSL). Extensive unvegetated flats at Tobacco Range, Twin Cays, and Blue Ground Range have been attributed to catastrophic damage by hurricanes (Woodroffe, 1995). However, the location of these dieback areas primarily in the interior of the islands is inconsistent with this hypothesis. An alternative explanation is that soil conditions affecting the growth of mangroves varies with distance from the shoreline and that infrequent flushing of the island interior by tides limits soil aeration and allows toxins to build up (Fig. 6). Variation in topographic relief, reflecting underlying coral ridges and depressions (Macintyre et al., this volume), may interact BS) with hydrology to generate these spatial differences in hydroedaphic conditions that ultimately influence mangrove establishment, growth, and survival (McKee, 1995). A major question not yet resolved concerns the relative importance of biogenic versus geomorphologic processes in mangrove development. Early ecologists introduced the idea of mangroves as "land builders" or "walking trees" (Davis, 1940) because they could extend seaward by the accumulation of sediment around the aerial root system, which would elevate the soil surface and in turn allow seaward colonization. Geomorphologists later argued that mangroves were just passive players in shoreline changes, simply responding to the physical processes of sedimentation, erosion, and changing sea level (Egler, 1952; Spackman et al., 1966; Thom, 1967; Wanless, 1974). Woodroffe (1983), however, pointed out the differences between geomorphologically active areas such as deltas and estuaries with large allochthonous inputs of sediment and carbonate settings in which peat deposition is the major sedimentary process. Woodroffe (1981) and Ellison (1993) have shown that mangroves in carbonate settings keep pace with rising sea level primarily through the deposition of peat. A similar conclusion may be made for the Pelican Cays in the Belize Barrier Reef complex. Cores obtained from the Tobacco Range Islands (Cameron and Palmer, 1995; Macintyre et al., 1995), Twin Cays, and Turneffe Atoll (McKee, unpublished data) show that other mangrove islands in Belize are similarly underlain by peat, which in some cases is up to 10 m in depth. Although deposits of sand or coral may be found at the base and as thin lenses within the peat core profiles (Cameron and Palmer, 1995; Macintyre et al., 1995; Fig. 8, this study; McKee, unpublished data), the bulk of the substrate underlying these islands is of mangrove origin. Surface trapping of calcareous sand by mangrove aerial roots occurs but is limited to island fringes and tidal creeks. This surface trapping may facilitate the seaward expansion of mangroves, but the development of an extensive below-ground root system is necessary to bind and consolidate the sand deposited along mangrove margins. Also, it is apparent from the Cat Cay cores that once mangroves establish, the main process maintaining surface elevations is peat formation. The high organic matter content of these cores provides further evidence that these systems are being maintained primarily by the deposition of organic matter. Emphasis on geologic rather than biogenic processes may create an erroneous impression of what controls the development and maintenance of healthy mangrove ecosystems in sediment- poor environments (Woodroffe, 1983). In particular, few appreciate the integral role that mangrove vegetation plays in the growth of islands in the Belize Barrier Reef complex (McField et al., 1996). The preliminary findings reported here illustrate the direct contribution of mangrove organic matter, particularly roots, to vertical accretion at Cat Cay. The increase in wood and leaf fragments in the upper 20 cm of peat cores from the dieback zone is consistent with observations of large quantities of standing and fallen leaf and wood litter at recent dieback sites elsewhere (McKee, personal observation). However, surface litter accounts for a relatively small proportion of mangrove peat overall. Unlike aerial litter, mangrove roots generally remain where they are produced and decompose extremely slowly owing to the anaerobic conditions in the sediment (McKee et al., 1988; McKee, 1995; Fig. 6, this study). Surface litter is also subject to tidal export as well as to consumption by animals such as crabs and snails (Robertson, 1991). Our findings thus indicate that root production and decomposition are two important processes controlling peat formation and hence vertical accretion and the ability of these islands to keep pace with rising sea level. 56 The origin of current vegetation patterns cannot be understood without some knowledge of past vegetation patterns and the manner in which mangroves have established and developed on these islands. Like Davis (1940), we found Rhizophora-derived peat beneath peat produced by Avicennia. Davis's interpretations, which were based primarily on macroscopic peat characteristics, were criticized by Cohen and Spackman (1977). Unlike Davis, however, we identified the peat from Cat Cay cores down to species level with the aid of both macroscopic and microscopic features. The success of this technique depends on careful and extensive examination of living and fossil plant material from different environments (Cohen and Spackman, 1977; McKee, unpublished data). Quantification of identified root fragments revealed that relative abundance of mangrove species varies over depth in the peat cores. In what is now a dieback area, red mangroves initially established upon a coral base and gradually built peat vertically. This colonization process can be observed occurring on shallow submerged reef ridges around Cat Cay and nearby islands, as noted by others (Macintyre et al., this volume). Avicennia appeared later, increasing in abundance and expanding outward from the island's center. At some point in the recent past, the center of the Avicennia stand began to deteriorate, creating a central dieback area devoid of emergent vegetation. Similar peat patterns of initial colonization by Rhizophora, followed by invasion, expansion, and dieback of Avicennia can be found at Twin Cays and Turneffe Atoll in the central and northern parts, respectively, of the Belize Barrier Reef system (McKee, unpublished data). Dieback areas on some of these islands are ultimately recolonized by Rhizophora (Woodroffe, 1995; McKee, personal observation), suggesting that dieback is part of a natural cycle of development and deterioration in response to changes in hydroedaphic conditions controlling dispersal, establishment, growth, and mortality of mangroves. Although the ultimate control over mangrove development and succession at Cat Cay is sea-level change, this external physical force affects in situ production and decomposition of mangrove organic matter, which in turn determine surface elevations relative to sea level. CONCLUSIONS In contrast to forests receiving large quantities of terrigenous or marine sediment (Woodroffe, 1983), the future existence of the Pelican Cays and other mangrove islands in the Belize Barrier Reef complex depends on the continued presence of mangroves and their contribution to peat formation. A lack of understanding and insufficient documentation of the direct contribution by mangroves to soil formation has many serious consequences, including policies allowing the trimming and clear-cutting of mangrove forests for urban expansion, tourist resorts, and shrimp ponds, to name a few activities occurring in the coastal zone of Belize (MacField et al., 1996). Because mangroves make a major contribution to soil formation, their removal eliminates a major source of material that maintains surface elevations in Belize's coastal zone. Since eustatic sea level rise is predicted to accelerate in the next century (Wigley and Raper, 1993), protection of mangroves will be essential to prevent submergence of these oceanic islands. The demise of these island forests would not only represent a significant land loss, but organisms unique to this habitat, particularly those solely dependent on mangroves for substrate, nursery, refuge, and food (Riitzler and Feller, 1996), would also disappear. Thus, although mangrove systems throughout Belize will be affected by rising sea level and other global changes, oceanic island forests such as the Pelican Cays will be the most vulnerable and 57 will require management plans based on a clear understanding of the role mangroves play in counterbalancing sea-level rise. ACKNOWLEDGMENTS Fieldwork for this project was supported by the National Museum of Natural History's Caribbean Coral Reef Ecosystem Program (CCRE Contribution No. 578). We thank Erik Chick, John O'Neil, Dennis Whigham, and Candy Feller for field assistance. Don Cahoon and Beth Vairin provided comments on the manuscript. REFERENCES Cameron, C. C., and C. A. Palmer 1995. The mangrove peat of the Tobacco Range islands, Belize Barrier Reef, Central America. Afoll Research Bulletin, no. 431, 32p. Cohen, A. D., and W. Spackman 1977. Phytogenic organic sediments and sedimentary environments in the Everglades- mangrove complex. Part II. The origin, description and classification of the peats of southern Florida. Palaeontographica B 162:71-114. Davis, J. H., Jr. 1940. The ecology and geologic role of mangroves in Florida, Papua Tortugas Laboratory 32:304-412. Carnegie Institute, Washington Publication no. 517. Eglenike cE. 1952. Southeast saline Everglades vegetation, Florida, and its management. Vegetatio 3:213-265. Ellison, J.C. 1993. Mangrove retreat with rising sea-level, Bermuda. Estuarine, Coastal and Shelf Science 37:75-87. Macintyre, I. G., M. M. Littler, and D. S. Littler 1995. Holocene history of Tobacco Range, Belize, Central America. Atoll Research Bulletin, no. 430, 18p. McField, M., J. Gibson, and S. Wells 1996. The State of the Coastal Zone Report. Fisheries Department, Belize City, Belize. McKee, K. L. 1995. Seedling recruitment patterns in a Belizean mangrove forest: Effects of establishment ability and physico-chemical factors. Oecologia 101:448-460. McKee, K. L., I. A. Mendelssohn, and M. W. Hester 1988. Reexamination of pore water sulfide concentrations and redox potentials near the aerial roots of Rhizophora mangle and Avicennia germinans. American Journal of Botany 5:1352-1359. Parent, L. E., and J. Caron 1993. Physical properties of organic soils. In Soil Sampling and Methods of Analysis, edited by M. R. Carter, 441-458. Boca Raton: Lewis Publishers. Parkinson, R. W., R. D. DeLaune, and J. R. White 1994. Holocene sea-level rise and the fate of mangrove forests within the wider Caribbean 58 region. Journal of Coastal Research 10:1077-1086. Robertson, A. I. 1991. Plant-animal interactions and the structure and function of mangrove forest ecosystems. Australian Journal of Ecology 16:433-443. Riitzler, K., and I. C. Feller 1996. Caribbean mangrove swamps. Scientific American 274:94-99. Spackman, W., C. P. Dolsen, and W. Riegel 1966. Phytogenic organic sediments and sedimentary environments in the Everglades- mangrove complex. Part 1. Evidence of a transgressing sea and its effect on environments of the Shark River area of southwest Florida. Palaeontographica B MFAZSS—152- Thom, B. G. 1967. Mangrove ecology and deltaic geomorphology, Tabasco, Mexico. Journal of Ecology 55:301-342. Wanless, H. R. 1974. Mangrove sedimentation in geologic perspective. In Environments of South Florida. Present and Past, edited by P. J. Gleason. Miami Geological Society Mem. No. 2. Miami, Fla. Wigley, T. M. L., and S. C. B. Raper 1993. Future changes in global mean temperature and sea level. In Climate and Sea Level Change: Observations, Projections, and Implications, edited by R. A. Warrick, E. M. Barrow, and T. M. L. Wigley, 11-33. Cambridge: Cambridge University Press. Woodroffe, C. D. 1981. Mangrove swamp stratigraphy and Holocene transgression, Grand Cayman Island, West Indies. Marine Geology 41:271-294. 1983. The development of mangroves from a geological perspective. In Biology and Ecology of Mangroves, edited by H. J. Teas, 1-17. The Hague: Dr. W. Junk. 1995. Mangrove vegetation of Tobacco Range and nearby mangrove ranges, central Belize barrier reef. Afoll Research Bulletin, no. 427, 35p. ATOLL RESEARCH BULLETIN NO. 469 PRELIMINARY HYDROGRAPHIC SURVEYS OF SOME PONDS IN THE PELICAN CAYS, BELIZE BY DANIEL W. URISH ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 Pelican Cays Northeast Cay Bird Coys; Co Cat Cay: f » & BB NS Ayicennia Cay @¢°3 Guatemala Honduras Lagoon Cays ® -#————— lkm Reine tetevafein/-/sevoiae eae reef SS Se tidal flat Figure 1. Map of Pelican Cays, Belize, Central America and pond identifications. PRELIMINARY HYDROGRAPHIC SURVEYS OF SOME PONDS IN THE PELICAN CAYS, BELIZE, CENTRAL AMERICA BY DANIEL W. URISH! ABSTRACT Hydrographic and hydrologic surveys of three ponds in the Pelican Cays Group were carried out using hand-held sonar, pressure transducer tide loggers, and YSI temperature-salinity probes. These ponds characteristically have a circular pattern with steeply sloping sides. The pond depths range from 4.6 m in a small pond with a diameter of 100 m to 15.2 m ina large pond with a diameter of 840 m; in general, depth increases with increasing pond size. Depths ranged from 15.2 m for a distance of 100 m between cays to 28.7 m for a distance of 1,150 m. Pond salinities averaged 35.3 ppt with a range of 1.5 ppt, and temperatures averaged 31.4° C with surface temperatures about | ° C higher than bottom temperatures. INTRODUCTION The Pelican Cays lie some 13 km east of the coast of Belize on the Belize Barrier Reef (Fig. 1). The submerged limestone shelf on which the cays are located have the characteristics of submerged karst limestone topography (Purdy, 1974; Stoddart et al., 1982). These cays are believed to have been formed on an ancient subaerially eroded limestone plateau that was submerged during the Holocene Transgression some 6,000—8,000 years B.P. According to Macintyre et al. (this volume), the circular pattern of the ponds in these cays is related to a combination of karst control and differential reef growth. The cays are composed of coral and coral rubble covered largely by mangrove forest. They can be characterized as low-lying land masses of coral and peat surrounding central ponds. The coastal margins of the cays are vegetated by the red mangrove Rhizophora mangle in the lower tidal washed sections, and some black mangrove, Avicennia germinans, in the slightly higher regions. Sand deposits with coconut trees can be found in a few higher areas. The ponds, which also appear as well-delineated circular bottom configurations in the outside lagoon, have steep walls. Many of the ponds overlap, separated only by articulate ridges. Some ponds are completely enclosed, but most are connected with other ponds by shallow channels. Tides along the barrier reef are microtidal and of a mixed semidiurnal nature, with a mean range of only about 15 cm, but this, along with windset, is sufficient to cause some interpond currents (Kjerfve et al., 1982). However, because most of the ponds have relatively deep central areas, they are probably poorly flushed, except in the upper water, and are essentially individual ‘Department of Civil and Environmental Engineering, University of Rhode Island, Bliss Hall, 1 Lippitt Road, Kingston, RI 02881. 62 ecosystems. There is no direct precipitation record for the region, but extrapolation from mainland records (Walker, 1973) and limited records at Carrie Bow Cay (Riitzler and Ferraris, 1982) suggest that annual rainfall averages about 190 cm/yr (75 in/yr). The nearest official weather station is Melinda Forest Station, located on the mainland coast about 45 km to the north. A dry period is common from February through May, when the rainfall averages only about one-third that of other months. Hurricanes during the months of July to October can produce storm surges that completely inundate the low-lying cays. During the periods of study, pond conditions were little affected by precipitation or unusual sea conditions. During the period May 18—26, 1994, we investigated the ponds of three cays of the Pelican Cays Group: Pond A (Cat Cay), Pond C (Manatee Cay), and Ponds E, F and G (Fisherman’s Cay) (Fig. 1). Additional depth measurements and investigations of the broader hydrographic characteristics of the regions between the cays were recorded during the period May 7-10, 1995. The hydrographic surveys were part of a larger scientific field study of the Pelican Cays undertaken by a group of scientists associated with the Smithsonian Institution. METHODS Very limited mapping information is available on the Pelican Cays. Some hydrographic mapping of the area can be found in U.S. naval charts based largely on British surveys between 1830 and 1841 (U.S. Navy, 1942). In order to obtain the scale of detail necessary for this study, the island-pond maps (Figs. 3, 6, and 10) were drawn using aerial photographs taken by Tony Rath (Dangriga, Belize) during May 1995. Distances and dimensions were ground-truthed in the field and checked against other published maps. Pond-bottom depth measurements were made using a Scuba-pro Model hand-held sonar deployed from a small boat. Survey line control across the ponds was established by making runs between prominent shoreline points. Locations of measurements along these lines were determined by "dead reckoning." That is, the measurements were made at regular timed intervals as the boat moved at a steady speed along the survey line. Location and depth in the interpond channels were determined by direct tape measurements. The latitude-longitude location shown on the maps was obtained with a Garmin GPS Model 50 unit. The tide record was obtained with the aid of pressure transducers and an Enviro-lab Model 120 data logger in the ponds. The data logger was programmed to collect level data at intervals of 15 min. The period of data collection was two and a half days. The vertical reference datum was arbitrary but was selected to keep all readings as positive elevations. Vertical profile depths were determined with direct tape measurements. Temperature and salinity values for the depth profiles were obtained using a YSI Model 33 instrument. Measurement of temperature and salinity were taken at 0.6-m (5.0-feet) intervals. RESULTS A two-and-one-half-day record of tidal fluctuation in Pond E at Fisherman’s Cay was obtained (Fig. 2). Pressure transducer probes used to measure water level were located in the mangrove roots on the west side of Pond E of Fisherman’s Cay as well as in Pond F. Although channel flow could be observed as the tide changed, there was very little measurable difference 63 in elevation or lag time. It is believed, however, that such differences can be measured using a shorter data collection interval, such as one minute. The location of work in Pond C at Cat Cay is shown on Fig. 3. It consisted of two bottom hydrologic profiles (Figs. 4 and 5), and two hydrologic depth profiles (Figs. 24 and 25) in the deeper parts of the ponds. The location of work in Pond A at Manatee Cay is shown on Fig. 6. It consisted of three hydrographic bottom profiles (Figs. 7—9) and two hydrographic depth profiles (Figs. 26 and 27) in the deeper parts of the pond. The field work at Fisherman’s Cay consisted of pond-bottom profiles and vertical hydrologic profiles in the deep parts of the ponds (Fig. 10). Four interpond channel cross sections (Figs. 11-14) were completed, two at the entrance into Pond E and two between Ponds E and F. Nine hydrographic lagoon bottom profiles (Figs. 15—23) and four hydrologic depth profiles (Figs. 28-31) were made. The locations of the 16 hydrographic survey runs made in May 1995 between the various cays of the Pelican Cays Group are shown on Fig. 32. Table 1 summarizes the results of these survey runs. Depths of water between cays ranged from 15.2 m for a distance of 100 m between cays to 28.7 m for a distance of 1,150 m. Table 1. Summary of pond hydrographic and hydrologic characteristics. Maximum Maximum Salinity Temperature Depth Length Depth Upper — Lower Upper Lower Location Profile (m) (m) (ppt) (ppt) C@) Ge) Cat Cay Pond A Al 250 Sy 3521 34.9 310 29.0 Manatee Cay Pond C Cl 130 14.0 357) 35.4 31.0 29.6 C2 200 10.7 35.6 3501 Sia) 30.0 Fisherman’s Cay Pond E Pond F Pond G sill 100 4.6 Boal 34.7 SEZ 31°) Fal 80 32 B50 36.2 31.0 30.0 64 Figure 2. Plot of tidal fluctuations in Pond E, Fisherman’s Cay during period of May 21- 23 1994. 65 LEGEND 1— 1” BOTTOM PROFILE > DEPTH PROFILE e GPS LOCATION @ 16°39.65'N ~ 88° 11.11'W POND A / 16° 39.44'N 88° 10.99' W Figure 3. Map of Cat Cay showing locations of bottom profiles and depth profiles for Pond A. 66 100 200 m DEPTH 60 ft 0 200 400 600 800 1000 DISTANCE FROM SOUTH EDGE Figure 4. Pond A - Cat Cay Bottom Profile 1-1". ‘5 50 100 m I ~ 10 5 20 = 30 0. 10 Q 40 15 50 60 ft 0 100 200 300 400 500 DISTANCE FROM WEST EDGE Figure 5. Pond A - Cat Cay Bottom Profile 2-2'. 16° 39.37'N 88° 11.53' W ; 67 LEGEND 3— 3” BOTTOM PROFILE x DEPTH PROFILE @ GPS LOCATION MANATEE CAY Figure 6. Map of Manatee Cay showing locations of bottom profiles and depth profiles for Pond C. 68 40 80 120 m DEPTH ft 0 100 200 300 400 500 DISTANCE FROM WEST EDGE Figure 7. Pond C - Manatee Cay Bottom Profile 3-3". ; 10 20 30 m 4 4’ 10 4 = 20 jeu Lu Q 8 30 40 ft 0 20 40 60 80 100 120 DISTANCE FROM EAST EDGE Figure 8. Pond C - Manatee Cay Bottom Profile 4-4". 100 200 DEPTH ft 0 200 400 600 800 DISTANCE FROM SOUTH EDGE Figure 9. Pond C - Manatee Cay Bottom Profile 5-5’. 300 1000 1200 69 70 LEGEND Q— 9” BOTTOM PROFILE 8’ x DEPTH PROFILE e GPS LOCATION POND E N 16° 40.23' ‘93 W 88° 11.39 © E2X haat oH 0 100m FISHERMANS CAY Figure 10. Map of Fisherman’s Cay showing locations of bottom profiles and depth profiles for Ponds E, F, and G. 71 —~a—- PONDE CROSS-SECTION Y Depth ft 0 10 20 30 40 Distance Figure 11. Fisherman’s Cay Channel Axial Profile from Pond E to lagoon. 0 2 4 6 8 m ‘ 0.5 2 cS a a 3 1.0 4 5 ft 0 10 20 30 Distance Figure 12. Fisherman’s Cay Cross-Section looking into Pond E. W2 CROSS-SECTION ft 0 10 20 30 40 50 Distance Figure 13. Fisherman’s Cay Channel Axial Profile from Pond F to Pond E 1 2 3 m 0 1 5) 2 £s fou ® 3 a (0; 4 5 ft 0 2 4 6 8 10 12 Distance Figure 14. Fisherman’s Cay Cross-Section looking from Pond E into Pond F M3 f 50 100 150 200 m 10 5 20 ao - Qa a 30 10 40 50 15 ft 0 200 400 600 800 DISTANCE FROM SOUTH EDGE Figure 15. Pond G - Fisherman’s Cay Bottom Profile 6-6. : 100 200 m 7 7 10 4 AG 620 LU Qa 8 30 40 ft 0 200 400 600 800 1000 DISTANCE FROM WEST EDGE Figure 16. Fisherman’s Cay Bottom Profile 7-7’ 74 DEPTH ft 0 100 200 300 400 DISTANCE FROM WEST EDGE Figure 17. Pond E - Fisherman’s Cay Bottom Profile 8-8' 20 40 m 0 5 2 = 10 a Lu QO 4 15 20 ft 0 100 200 DISTANCE FOM NORTH EDGE Figure 18. Pond E - Fisherman’s Cay Bottom Profile 9-9". 10 10 DEPTH 20 ft 0 100 200 300 DISTANCE FROM NORTH EDGE Figure 19. Pond F - Fisherman’s Cay Bottom Profile 10-10". 100 25 50 75 (oe) DEPTH ft 0 100 200 DISTANCE FROM WEST EDGE Figure 20. Pond F - Fisherman’s Cay Bottom Profile 11-11". 400 75 76 6 20 40. m 2 12')2 8 10 “ 12 W 4 14 16 18 ft 0 100 200 DISTANCE FROM WEST EDGE Figure 21. Pond F - Fisherman’s Cay Bottom Profile 12-12". 4 20 40 m 13 13" 6 2 8 = ae [ou Lu Q 12 4 14 16 ft 0 100 200 DISTANCE FROM EAST EDGE Figure 22. Pond F - Fisherman’s Cay Bottom Profile 13-13' 3 10 20 m 14 14 10 4 20 = 8 ui 30 fa) 12 40 50 ft 0 20 40 60 80 DISTANCE FROM EAST EDGE Figure 23. Fisherman’s Cay Bottom Profile 14-14' outside of Pond E. HG 78 ft TEMPERATURE (°C) o3 TEMPERATURE 1 SALINITY DEPTH 28 29 430 ~~ 31 32 33: #34 35 "36 SALINITY (ppt) Figure 24. Pond A - Cat Cay Depth Profile Al. TEMPERATURE (°C) NS TEMPERATURE SALINITY 10 20 Depth 30 40 7 29 30 31 32 33 34 35 36 SALINITY (ppt) Figure 25. Pond A - Cat Cay Depth Profile A2. 79 TEMPERATURE (°C) on 10 TEMPERATURE SALINITY gy x 20 a = hr x a 30 a 40 50 ; 238-29) 30 31 $2.33 (34-35 «3637 SALINITY (ppt) Figure 26. Pond C - Manatee Cay Depth Profile C1. ft TEMPERATURE (°C) m 0 0. | TEMPERATURE | 10 sanity | = 20 = Oo Lu Q 30 40 : 2829 30 31 32 33 34 35 36 §=637 38 SALINITY (ppt) Figure 27. Pond C - Manatee Cay Depth Profile C2. 80 TEMPERATURE (°C) ft 0 2 ® LS TEMPERATURE SALINITY 7 DEPTH 10 V2 28 30 32 34 SALINITY (ppt) Figure 28. Pond E - Fisherman’s Cay Depth Profile El. TEMPERATURE (°C) SALINITY TEMPERATURE 10 Depth 28 30 32 34 SALINITY (ppt) Figure 29. Pond F - Fisherman’s Cay Depth Profile F1. 36 o3 36 o3 38 TEMPERATURE (°C) ~~ TEMPERATURE SALINITY DEPTH 28 30 32 34 SALINITY (ppt) Figure 30. Pond G - Fisherman’s Cay Depth Profile G1. ft TEMPERATURE (°C) 20 TEMPERATURE 30 SALINITY Depth 40 50 60 29 30 31 32 33 SALINITY (ppt) Figure 31. Fisherman’s Cay Depth Profile E2. 36 34 35 81 82 SOOO Northeast Cay Pelican Cays V/ f x SZ) Ridge Cay Co Cat Cay: :.. Manatee\Cay S 1Ge oe Little Cat = eames survey run if: island numerous coral heads reef no, ------ tidal flat ae = Boe 29m Figure 32. Map of Pelican Cays showing locations of inter-cay hydrographic survey runs. 83 DISCUSSION Pond depths ranged from 4.6 m (15 feet) in relatively small Pond E at Fisherman’s Cay to 15.2 m (50 feet) in the large Pond A at Cat Cay. In general, there appears to be a direct correlation between size and depth in the ponds. Such a correlation might be better determined by more detailed examination of the size of the circular bottom contour pattern and the depth of the configuration. The large ponds seem to be a composite of smaller circular depressions. Results for the ponds are summarized in Table 2. Table 2. Inter-cay water depths of Pelican Cays group. Survey run locations are shown on Figure 32. Survey Run Distance (m) Max. Depth (m) The variation in salinity within ponds, or between ponds is noteworthy, but not remarkable. The near-surface water salinity was 35.1 ppt, except in Manatee Cay, which averaged 35.6 ppt, while the bottom water was in most cases about 0.2—0.5 ppt or less. In two ponds at Fisherman's Cay, the bottom salinity was slightly higher. Near-surface salinities are likely to be significantly less during the rainy season. The variation in temperature within ponds and between ponds was not large but is consistent and significant. Near-surface water temperatures ranged from 31.0° C to 31.8° C, averaging 31.3° C, while bottom water temperatures ranged from 29.0° C for the deepest reading in Pond A at Cat Cay to 31.5° C for the shallowest reading in Pond E at Fisherman's Cay. This is not surprising since water temperature is a result of solar radiation and the volume of water available to absorb it. The measurements were all taken during the middle of generally sunny 84 days. Significant semidiurnal, as well as seasonal, variations in the near-surface temperature are to be expected. Analysis of the hydrographic survey data for the lagoon regions between the cays (Table 2) shows that the sides of the cays exhibit very steep slopes similar to the ponds and relatively flat bottoms over the major length of the runs. Analysis of 23 lagoon slopes gives an average of 28°, considerably greater than the 17° of the ponds. Again, there appears to be a direct correlation between water depth and distance across open water, whether in the ponds of the cays or in the lagoon between cays. CONCLUSIONS Hydrographic and hydrologic surveys of ponds in three cays—Cat Cay, Manatee Cay, and Fisherman’s Cay—of the Pelican Cays Group show that the pond bathymetry characteristically shows a circular pattern with steeply sloping sides. An articulated ridge frequently occurs between these circular patterns. The pond depths range from 4.6 m in a small pond, such as Pond E in Fisherman’s Cay (100 m wide) to 15.2 m ina large pond such as Pond A in Cat Cay (250 m wide). The average maximum pond depth is 10.7 m. In general, depth tends to increase with the width of the pond. Measurenents of water depth between the cays showed maximum depth ranged from 15.2 m over a survey run distance of 100 m to a maximum depth of 28.7 m over a survey run distance of 1,150 m. The average maximum depth of the intercay survey runs is 22.7 m, which is more than twice that of the interior ponds. Here, depth tends to increase with increasing distance between cays. Side slopes in both ponds and the lagoon are steep, averaging 17° for the ponds and 28° for the lagoon. Maximum salinities in the ponds averaged 35.3 ppt near the surface and 35.2 ppt near the bottom, with a range of 1.5 ppt. There was a slight tendency for salinity to be lower near the bottom. Temperature showed a somewhat larger variation, ranging from an average maximum of 31.3° C near the surface to 30.2° C near the bottom, with a relative strong tendency for surface temperatures to be about 1.0° C higher at the surface. ACKNOWLEDGMENTS Special thanks are given to Carrie Bow Cay station managers Tom Hawkins (1994) and Steve Hays (1995) for their valuable assistance in the field operations of this study, and to Tony Rath of the Pelican Beach Hotel for his very useful aerial photographs of the cays. Field support for this project was provided by the National Museum of Natural History’s Caribbean Coral Reef Ecosystem Program (CCRE Contribution No. 579). REFERENCES Kjertve, B., K. Riitzler, and G. H. Kierspe 1982. Tides at Carrie Bow Cay, Belize, in the Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize. In The Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize, 1. Structure and Communities, edited by Klaus Riitzler and Ian G. Macintyre. Washington, DC: Smithsonian Institution Press. 85 Purdy, EG: 1974. Karst-determined facies patterns in British Honduras: Holocene carbonate sedimentation model. American Association of Petroleum Geologists 58:825-855. Purdy, E. G., and G. T. Bertram 1993. Carbonate Concepts from the Maldives, Indian Ocean. American Association of Petroleum Geologists Studies in Geology, No. 34, 56 p. Riizler, K., and J. D. Ferrais 1982. Terrestrial Environment and Climate, Carrie Bow Cay, Belize. In The Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize, |. Structure and Communities, edited by Klaus Riitzler and Ian G. Macintyre. Washington, DC: Smithsonian Institution Press. Stoddart, D. R., F. R. Fosberg, and D. L. Spellman 1982. Cays of the Belize Barrier Reef and lagoon (western Caribbean Sea). Atoll Research Bulletin 256:1-76 . U.S. Navy 1942. Oceanographic Office Map of British Honduras: Ranguana Cay to Columbus Cay. Walker, S. H. 1973. Summary of Climatic Records for Belize, Land Resources Division, Surbiton, Surrey, England, Supplement No. 3. Rainfall distribution map contained in Belize, Country Environmental Profile, Robert Nicolait & Associates, USAID Contract No. 505-0000- C-00-3001-00, April 1994. 86 ATOLL RESEARCH BULLETIN NO. 470 HYDROGRAPHY OF A SEMI-ENCLOSED MANGROVE LAGOON, MANATEE CAY, BELIZE BY TRACY A. VILLAREAL, STEVE L. MORTON, AND GEORGE B. GARDNER ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 : Z Caribbean y 2 tee Belize City Dangriga Channel Guatemala Honduras Figure 1. Manatee Cay Pond C sample site (asterisk). HYDROGRAPHY OF A SEMI-ENCLOSED MANGROVE POND, MANATEE CAY, BELIZE BY TRACY A. VILLAREAL,' STEVE L. MORTON,” AND GEORGE B. GARDNER? ABSTRACT Hydrographic surveys using a recording CTD system in 1995 and 1996 indicate that the semi-enclosed pond C in Manatee Cay has a distinct hydrographic, chemical, and biological signal. Pond water appears to be channel water advected into the pond surface and modified by local heating and evaporation. Temperature-salinity relationships suggest episodic formation of pond water and a long residence time, although the time scale for this is not known. Pond phytoplankton populations in these two years were dominated by migrating dinoflagellate populations (Gymnodinium sanguinium), although there is evidence that periodic diatom blooms created a midwater silicate minimum and near-bottom maximum as a result of midwater uptake and near-bottom remineralization. INTRODUCTION The Belize Barrier Reef is the second largest reef system in the world, and the largest in the Western Hemisphere (Riitzler and Macintyre, 1982). It extends approximately 250 km from the Yucatan Peninsula to the Gulf of Honduras (James and Ginsburg, 1979). North of Belize City, the shelf is shallow and has a series of islands with a discontinuous reef lacking a well- defined reef flat. South of Belize City, there is a well-developed barrier platform that averages 4 to 5 m deep (Stoddart et al., 1982). In the southern reaches of the platform near the latitude of Gladen Spit, the barrier reef is cut by deep channels that form a number of shelf atolls, or faros (James and Ginsburg, 1979). The Pelican Cays group is located near this region where shallow mangrove cays are immediately adjacent to deep channels on the order of 20 to 30 m in depth. Cay morphology is unusual in that several of the group have relatively deep central ponds separated by shallow sills from the adjacent channels. This honeycomb pattern originates in the topography of the underlying Pleistocene karst relief (Macintyre, this volume). The lagoon-like ponds may be 10 to 12 m deep and harbor rich tunicate and sponge populations on the fringing ' Marine Science Institute, The University of Texas at Austin, 750 Channel View Dr., Port Aransas, TX 78373. "NOAA National Ocean Service, Center for Coastal and Environmental Health and Biomolecular Research, 219 Fort Johnson Road, Charleston, SC 29412. Environmental, Coastal and Ocean Sciences Program, University of Massachusetts, 100 Morrissey Blvd., Boston, MA 02155. 90 mangrove prop roots (Goodbody, this volume, Riitzler et al., this volume). Within the ponds, the encrusting prop root fauna and the mangroves have complex interactions and nutrient exchanges (Ellison et al., 1996, Riitzler and Feller, 1996). The importance of local processes is enhanced by the limited exchange with adjacent channel water resulting from only 30-cm tides and wind- driven circulation. Little is known of the phytoplankton dynamics in these ponds. The composition of the pond phytoplankton species is distinct from that found in the deep channels only a few meters away and is undoubtedly a result of the shallow, fjord-like sill that isolates most of the pond. Shallow cays to the north have been reported to contain intense blooms of dinoflagellates that visibly discolor the water red (Morton and Villareal, in press). However, there are no quantitative surveys of the phytoplankton in the Pelican Cays ponds, and only limited hydrographic surveys (Urish, this volume). We report here on the general characteristics of one of these ponds (C) in Manatee Cay and briefly describe the phytoplankton community during late May for two years. METHODS Surveys were conducted from small boats during May 1995 and 1996 in the north pond (C), Manatee Cay (16°40.07'N, 88°09.64'W). The general sample site and location are illustrated in Fig. 1. Time-series sampling was conducted in the center of the pond while the boat was anchored. For comparison, casts were also taken outside the pond in the channel and at an open ocean station approximately | km east of Carrie Bow Cay (16°48'N, 88°05'W). Hydrographic data (salinity, temperature, depth) were collected in 1996 using a Sea Bird Model 19 Seacat recording CTD equipped with a Sea Tech fluorometer. Data were logged internally at 0.5-sec intervals and downloaded at the field station at Carrie Bow Cay. Discrete chlorophyll, phytoplankton, and nutrient samples were collected in the pond at 1-m depth intervals using a hand pump in 1995 and a bottle sampler in 1996. Two discrete samples were collected by snorklers under the mangrove prop root fringe surrounding the pond at a depth of approximately 0.5 m. Secchi disk depth was determined immediately after the discrete water sample cast. Chlorophyll samples were returned to the lab (approximately 1 h) and filtered through Poretics GF-75 filters for total chlorophyll (both years) and on Poretics 5-um pore size polycarbonate filters for the > 5-zm fraction (1995 only). Chlorophyll was extracted in methanol overnight and measured either on a Turner Model 10-AU fluorometer using a nonacidification technique in 1995 (Welschmeyer, 1994) or on Turner Model 111 fluorometer in 1996 using an acidification technique (Parsons et al., 1984). Both fluorometers were calibrated to pure chlorophyll a. The 1- L phytoplankton samples were concentrated using a S5-um mesh net and fixed with 1% glutaraldehyde. Phytoplankton populations were enumerated using a 1-ml Sedgwick Rafter counting chamber. Nutrient samples were immediately filtered and placed on ice packs until their return to Carrie Bow Cay (approximately 1 h) where 15-ml samples were frozen and returned to the University of Massachusetts for analysis on a TrAACs 800 automated nutrient analysis system (nitrate + nitrite, silicate, phosphate, and ammonium). A separate 30-ml aliquot was measured on site for phosphate using a 10-cm spectrophotometer cell (Parsons et al., 1984). For purposes of clarity, pond stations refer to Manatee Pond C, channel stations refer to locations near the cays but in deep water (> 15 m), and ocean stations refers to the stations east of Carrie Bow Cay. Oy RESULTS Discrete chlorophyll samples in 1995 (Fig. 2) indicated a midwater maximum (5.8 wg L") on 5/23/95 (Fig. 2b) and a surface maximum of 5.0 wg L*' on 5/25/95 (Fig. 2c). A midwater maximum is suggested in the 5/21/95 sample (Fig. 2a); however, our initial sampling at three depths was inadequate to capture the detail found at 1-m resolution. Significant variability was found in the size structure. The > 5.0-m size fraction accounted for < 50% of the total chlorophyll a on 5/21/95 (Fig. 2a), but greater than 80% at the surface on 5/25/95 (Fig. 2c). Regardless of the day, the > 5.0-um fraction accounted for 25% of the total chlorophyll a and generally > 30-40% of the total. A Gymnodinium sanguinium population corresponded to the > 5-um chlorophyll a data for each date in 1995 (Fig. 3). The hydrographic casts from Manatee Cay in 1996 revealed a complex vertical structure that varied on a diurnal basis. Three time series were collected: 5/20/96, 5/23/96, and 5/25/96 (Figs. 4, 5, and 6, respectively). Several generalities are evident. Highest temperatures were found at the surface later in the day. Significant heating occurred during the day, as much as 0.45° C on 5/23/96 (Fig. 5a). Both 5/20/96 (Fig. 4a) and 5/23/96 (Fig. 5a) showed a complex temperature curve with a subsurface minimum increasing to a maximum and then decreasing to the bottom. The 5/25/96 data did not show this as clearly, although there was a slight inflection at approximately 6 m (Fig. 6a). Salinity showed no consistent pattern between the days. On 5/20/96, there was a complex structure with midwater minimum and maximum (Fig. 4b), while on 5/23/96 salinity increased to a sharp inflection at 6 m (Fig. 5b) and gradually increased with depth. On 5/25/96, salinity was similar throughout the upper 5 m and was separated from the more saline bottom water by a very sharp halocline at 5 m (Fig. 6b). The density structure on 5/20/96 was highly unusual (Fig. 4d). The observed T-S structure created a high density layer at 3.5 to 4.0 m that was 0.2 sigma-t units higher than the underlying layer. There was a small indication of a similar phenomenon on 5/23/96 (Fig. 5d), but no evidence of it on 5/25/96 (Fig. 6d). Fluorescence patterns were very similar on all days. There was < 1 fluorescent unit at the surface, which increased to a bottom or near-bottom maximum. The magnitude of the maximum varied from day-to-day, reaching a maximum of 12 units on 5/25/96 (Fig. 6c). The maximum on 5/23/96 (Fig. 5c) was clearly several meters above the bottom. Discrete chlorophyll values from bottle measurements (Fig. 7a) corresponded well with the fluorescence data in 1996 (r°= 0.92). On 5/23/96, values gradually increased from 0.5 wg L"! at the surface to 1.0 wg L’' at 7 m. Below that depth, they increased sharply to 11.5 wg chl Lat the bottom, where Gymnodinium sanguinium was again the dominant phytoplankton. Discrete nutrient samples (Fig. 7b) consistently showed a pattern of elevated concentration at the surface, a midwater column minimum, and increased concentrations in the lower 3 m. Individual nutrients did not track each other but tended to have slightly different patterns. Nitrate values were maximal at the surface (0.35 4M) and were elevated in the upper 3 m (0.20—0.28 4M), decreased to < 0.03 to 0.08 uM from 4 to 8 m and then increased by > 0.30 uM from 8 to 9 m to a bottom maximum of 0.35 4M from 9 to 11 m. Phosphate concentrations (0-1 m) were < 0.02 uM to 0.06 uM in the upper 2 m, were undetectable at 4-5 m, and then steadily increased to 0.10 uM below 6 m. This pattern was seen in both the manual (reported) and automated analyses. Ammonium was at detection limits in the upper 8 m, and then increased to a bottom maximum of Depth (m) Depth (m) Depth (m) Fig. 2. Size-fractionated chlorophyll measurements in Pond C from 1995. A. 5/21/95 at 11:45 h, B. 5/23/95: at 12:49 hy C. 5/25/95 at 15:15 h: 93 Cells L7! 0 500 1000 1500 2000 2500 3000 3500 4000 Depth (im) = fe 65/21/95 —@— 5/23/95 - A- 5/25/95 12 Fig. 3. Gymnodinium sanguinium cell abundance in Pond C in May 1995. Samples are from the same aliquots as the chlorophyl! data in Fig. 2. 94 c&e ‘Sp ONS, ‘Aysuaq °q ‘aousdsasonyy [[AydosopyD *— ‘ounyesadwia |, “gq “AWUI[ES “VW °96/0Z/S WO s]JUStOINSeaLU ULUNJOD J9}VM D puog Avo sayeuryy ‘p “S1y ‘dG vi (lu) ydeq Scammaces Seo Poe odd }-eWBIS AIS ISIE MONE Sy 1S} 7S) ay i re 1E ) gousosalon|4 96/0¢/S WV 00:00:21 96/0¢/S WY 00:06:11 96/0¢/S & WW 00:00:11 9 s 96/02/S 9 3 WY 00:08:01 p Z (0) e9e Loe 6SE LSE Sse pOE ZOE OE 8 6c 962 62 762 62 B8z (NSd) Ayunes 9) , einjeredwia | (w) udaq (w) ydaq 95 SEE ‘yjdop sues ou} ye JayeM ouURYO Jo AjIsuap SoyeoIpUl MOLY “(]-BWISIS) AyIsuaq ‘q ‘souadsasON yy [[AYydoso[ YD *|_ ‘ainjesodua |, ‘g ‘AWTS “V °96/€7/S WO S}UOWOINSvOU UUINJOS J9}eM D puod ABD so}euRYY “¢ “BLY ‘q = -> on wage Lay Vv M we) 2 “A -OL g = ao SI ~9 = EG ~0 CISC eC CumaG cca Siccm ico a .cice AMON Gh ae Gee GS Dag }-ewbis eousoselon|- 96/E¢/S WY OO: FP: ht ‘ Vv 96/EC/S | bl WY 00:08:01 ll 96/E/G | o WY 00:00:01 OL UO 3 Si 96/E7/G -8 3 | Wv 00:0€:60 ue 96/€2/S Wy 00:00:60 ™ Y =¢ ege 496 666 266 GE j 50 Oe Bad OG Vid Ale 62 (ASd) Aywuyes Do sunyesadua | (w) yydaq (Ww) ujdaq 96 ‘(Q-ewisis) Aysuaq “q ‘aouadsaiony [[AydosopyD “5 ‘oimyesodua | ‘q “AUTRES “VY °C6/SZ/S Woy s}UsWaINSvatH ULUNTOS Ja}eM J puog kvD sayeury 9 “SI ad fpr CO ole U3 a [oe a 9 ee ae [eee lic 7 (i T =) c&e €¢ Bide “Sigg ved dice }-ewbIS COCmEECOGm IGG GiGe: “sse oce (ASd) Ajuyes (tu) yydaq (wi) ujdeq Pee ae a eae al ee a ee pea yO) Grobe Ol 6G 8 2 OS) oy Cid) souso0saJon|4j Le/SZ/S ‘Vorb WY 00:20:01 eS: Le/Sz/S oe WY 00:51:60 ; 0 Os ed a® 3) «a = 9 {7 fo) = 0? a Lae ae =()) vOt cOE OF 86c 96¢C PE¢e 26¢e 62 9. ainyesadwa] (w) uideq (w) udaq Depth (m) uM NOx” (), PO4* (#) 6) 0.1 0.2 0.3 0.4 OFS 5 B. Pee iyA E — Ss 6 ® Q oy, os een ee ee SN am oe i ~@ ©) Oa Be ey ean pyle i 7shisy Ma uM Silicate (@), NH4* (¥) Fig. 7. Discrete bottle measurements from Manatee Cay Pond C on 5/23/96. A. Extracted chlorophyll a, B. Nutrients. 98 0.2 uM. Silicate decreased gradually from 0 to 4 m, had a pronounced midwater minimum at 7 m (1.90 uM), and then increased to its maximum (4.75—5.00 4M) in the lower 3 m. The surface Si values are similar to values noted previously from surface water at the offshore station and the coastal pond (Villareal, unpublished observation). Nitrate concentrations from under the prop root fringe were 0.7 and 1.1 4M in samples collected 2 m apart. Si was elevated in one sample (5.5 uM) while similar to the station water at the other. P was not detectable using the automated analysis (< 0.1 uM). The channel site (approximately 50 m outside the pond sill) had a surface temperature maximum (28.95° C) with a small subsurface temperature minimum at 5 m (Fig. 8). This was similar to that seen in the pond except that it was approximately 0.5 degrees colder and was 2 m deeper. Salinity increased uniformly with depth from 35.88 to 36.14 at 14.5 m, where it became nearly uniform from there to the bottom. Sigma-t increased from the surface to a sharp pycnocline at 14.5 m and was nearly uniform to the bottom. Fluorescence was low (< 0.2 units) and uniform throughout the water column and near the instrumental limit because of the scale setting (data not shown). Temperature-salinity relationships for the pond, channel, and offshore regions (Fig. 9) suggest that Manatee pond C has extremely limited exchange with the deeper water outside the sill. Pond water was consistently warmer and saltier than water at equivalent depths in the channel. Considerable daily variation was noted, with a steady progression of increased salinity and warmer temperatures from May 19 to May 23 (Fig. 10). This was particularly evident in the water below 36.3 PSU, which increased in temperature from 29.0 to 29.6° C. The sharp discontinuities noted in the lower 4—5 m of the pond suggest that this water does not originate in channel water overwashing the sill and sinking. This can be seen in the boxed area in Fig. 10, which represents the water below 6 m. Although the density (sigma-t) values for the channel and pond water intersect at this point (Fig. 5d), the channel water is 0.4 PSU lower and > 0.9° C cooler than the pond water at this depth. Pond water at this depth must be heavily modified by local heating and evaporative processes. The rapid changes from 5/20/96 to 5/25/96 above 6 m suggest a time scale of a few days or less for replacement or modification. DISCUSSION Hydrographic conditions in Manatee Pond are complex and appear to respond rapidly to local heating. The deep water of the pond appears to be hydrographically isolated from the surface, as is evident from both density profiles and nutrient distribution. The time scale for renewal is unclear. Limited fetch and intense heating create strong barriers to vertical mixing. The nutrient data suggest that the upper 3-4 m is modified channel water, as indicated by the Si concentrations (similar to surface channel water) and the salinity. The T-S profiles (Figs. 9, 10) also indicate that the pond water is not simply channel water but has been modified by local heating and evaporative process. We suggest that a thin layer of surface water advects into the pond, possibly by tidal or wind-driven circulation, and then is modified by local heating and evaporation to create the surface pond water. The rapid change in water above 6 m appears to occur within a few days. It is likely, although unknown at this time, that cooling increases density sufficiently to 99 ‘SBuIIAS JUSWUNISUI MOTOq SEM OUddSaION],J “(1-BUISIS) AyIsuaq “> ‘omnyesroduia |g ‘AyUITeS “VW °96/EZ/S Wos SJUSWINseaU UUINJOD JoyeM JoUURYD °g “OINST Gece ©a Sica icc atsce ice }eWBIS SS (15 Cis Ee OS ENS) SSIS. 6c 8°8¢ 9'8¢ 0 '8¢ (ASd) Ayuies 4 , ainyesadua | 100 36.6 Ocean clo ,otation Ne, meee ee > Zee © (ep) 2 Q e E=ECIOn = R50) ep) 35.8 35.6 Channel Stations Temperature ° C Fig. 9. Temperature-salinity plot including all 1996 ocean, channel and Manatee Cay Pond C stations. The boxes generally circumscribe the data points from a particular area, although a few outliers are evident. 101 . 36.6 36.4 So (ep) Qo P— i= 36 ow (ep) 35.8 35.6 28.8 Ee) 2S, 2G 29.6 29.8 380 30.2 30.4 Temperature °C Fig. 10. Temperature-salinity plot of the 1996 Manatee Cay Pond C data. Boxed area indicates deep water pool below 6 m. 102 cause sinking of this saline surface water. This is the only plausible mechanism that can create the high-salinity bottom water noted in the pond; however, the time scale for this process is unclear. The T-S relationships with the channel (Fig. 8) indicate that the bottom water is modified only slightly by daily heating and cooling, with most of the changes occurring in the upper 6 m. Density data (Fig. 5d) indicate that surface channel water would sink only to mid- depth and lacks the requisite T-S structure. Water below 6-m depth appears stable and isolated from the surface. In the absence of tropical storms or hurricanes, and excluding the possibility of cross-platform water infiltration, the deeper water mass may remain isolated on a seasonal basis. The most anomalous observation is the persistent density inversion seen clearly on 5/23/96. This distribution would have been unstable and not likely to persist for more than a few minutes. We have eliminated instrument error as a cause but are at a loss to explain the feature. Possibly, the complex organics released by the mangroves are modifying the equation of the state of the water or the relationship between conductivity and salinity. We are continuing to examine this anomaly. The mid-depth nutrient depletion noted on May 23, 1996, is clearly the result of biological activity. The near-bottom chlorophyll increase occurs at the same depth as the near- bottom nutrient increase and suggests that the phytoplankton were responding to the low-light, high-nutrient conditions at the bottom. In contrast to all the 1996 data, the 1995 data note midwater chlorophyll maxima on successive days. Unfortunately, we have no nutrient data from that year, and we are unable to determine conditions at that time. In 1995 most of the chlorophyll was in the larger size fraction, an unusual occurrence for oligotrophic waters. The 1996 midwater minimum and bottom maximum suggest that at some time scale, phytoplankton production consumes the nutrients, sinks or migrates, and is remineralized at depth. Possibly, diatom growth in the midwater strips out the Si and deposits it at depth as diatoms are remineralized. Previous surveys have noted extensive diatom blooms within these ponds, although we did not note them in this study (Faust, personal communication). The combined data from the two years suggest that phytoplankton populations are dynamic. Secchi disc depths indicate that the euphotic zone reaches the bottom and can support the 1996 bottom chlorophyll maximum. From the 1995 results, it appears that motile populations can create a midwater chlorophyll maximum. At some point, diatom growth stripped out the silicate in the mid-depth region and created an unusual bimodal nutrient distribution. Dinoflagellates are also dominant in the shallow pond of nearby Douglas Cay, the site of a red tide of Gonyaulax polygramma (Morton and Villareal, in press). These data provide only a snapshot of conditions in Manatee Cay and no information about rates. The lack of a phytoplankton peak at the surface where nutrients are abundant may be the result of photoinhibition or intense grazing by the mangrove prop root fouling community. The CTD data suggest that distinctive water masses occur in the pond, but we cannot identify any time scale for either overturn or flushing in the pond. Although we cannot estimate rates, the elevated biomass in the pond does raise questions about nutrient supply required to sustain this biomass. Tidal exchange is extremely limited in the pond, and the shallow sill suggests wind- driven circulation will be restricted to a meter or two. The dense fouling community on the prop roots consists primarily of sponges and tunicates that may aid remineralization by grazing or symbiotic nitrogen fixation, but this input would be expected to occur as ammonium, not nitrate. The elevated nitrate concentration under the prop root fringe suggests another source, either 103 symbiotic nitrification (Corredor et al., 1988) or processes in the mangroves. Phosphate was extremely depleted in the water column, and was consistent with possible P limitation in the mangroves themselves (Feller, 1995). Future work must include water column rate measurements, time-series hydrography, and mangrove-pond interactions in order to interpret the phytoplankton dynamics in the pond. ACKNOWLEDGMENTS Contribution No.580 of the Smithsonian Institution Caribbean Coral Reef Ecosystem Program. Contribution No. 1087 of the University of Texas at Austin Marine Science Institute. REFERENCES Corredor, J. E., C. R. Wilkinson, V. P. Vicente, J. M. Morell, and E. Otero 1988. Nitrate release by Caribbean reef sponges: Limnology and Oceanography 33:114-120. Ellison, A. M., E. J. Farnsworth, E. J., and R. R. Twilley 1996. Facultative mutualism between red mangroves and root-fouling sponges in Belizean mangroves. Ecology 77:243 1—2444. Feller, I. C. 1995. Effects of nutrient enrichment on growth and herbivory of dwarf red mangrove (Rhizophora mangle). Ecological Monographs 65:477-S0S. James, N. P., and R. N. Ginsburg 1979. The Seaward Margin of Belize Barrier and Atoll Reefs: Morphology, Sedimentology, Organism Distribution and Late Quaternary History. International Association of Sedimentologists, Special Publications No. 3, 1-191. Morton, S. L., and T. A. Villareal In press. An unusual red tide in a mangrove lagoon, Douglas Cay, Belize. Bulletin of Marine Science. Parsons, T. R., Y. Maita, Y., and G. M. Lalli 1984. A Manual of Chemical and Biological Methods for Seawater Analysis. New Y ork: Pergamon Press, 173 pp. Riitzler, K., and I. C. Feller 1996. Caribbean mangrove swamps. Scientific American 274:94-99, Riitzler, K., and I. G. Macintyre (eds.) 1982. The habitat distributions and community structure of the barrier reef complex at Carrie Bow Cay, Belize. In The Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize, I. Structure and Communities, edited by K. Riitzler and I. G. Macintyre, 9-46. Washington, DC: Smithsonian Institution Press. Stoddart, D. R., F. R. Fosberg, and D. L. Spellman 1982. Cays of the Belize Barrier Reef and lagoon (western Caribbean Sea). Atoll Research Bulletin 256: 1-276. Welschmeyer, N. A. 1994. Fluorometric analysis of chlorophyll a in the presence of chlorophyll 5 and pheopigments. Limnology and Oceanography 39:1985—-1992. ATOLL RESEARCH BULLETIN NO. 471 COMMUNITY STRUCTURE, WATER COLUMN NUTRIENTS, AND WATER FLOW IN TWO PELICAN CAYS PONDS, BELIZE BY THOMAS A. SHYKA AND KENNETH P. SEBENS ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 ay I Caribbean Guatemala Honduras ‘ 88 W aay a a | ae G ae, So i aN \-y / fe = aM P i 4 \ Ye ~ / \ e \ 0 5O 100m i ~S Pond C ‘ 1 / N —)) ip) // j Xf ee De : ee ! ey >) aay\ \ (on / ys SS / vA \\ | i] a ISS AS ie (32 as i . ~ e (=—— 4] Pond A 4/ ase ) Ca \ ' ! yA) \ \ DG — pau) eo } ( \ * yf i Mangrove XN ese | Nutrient Sampling Sites Cat Cay ae Manatee Cay Figure 1. Map showing location of nutrient sampling sites and Ponds A and C in Cat and Manatee Cay, respectively COMMUNITY STRUCTURE, WATER COLUMN NUTRIENTS AND WATER FLOW IN TWO PELICAN CAYS PONDS, BELIZE BY THOMAS A. SHYKA' AND KENNETH P. SEBENS? ABSTRACT Nutrient concentrations, water flow, temperature, chlorophyll a, and benthic community structure were measured at locations in two mangrove-ringed ponds in the Pelican Cays, Belize. Benthic community structure changed from coral dominance outside the ponds to macroalgae and seagrass dominance inside the ponds. While corals were generally absent inside Pond C, Manatee Cay, they persisted well into Pond A, Cat Cay. Dissolved inorganic nitrogen concentrations (ammonium + nitrate + nitrite) were naturally elevated along the coral ridges at the openings of the ponds and, sometimes, at the bottom of the ponds. Mean nitrate + nitrite concentrations, which ranged from 0.51 to 0.71 uM along the ridges at the entrances of both ponds between 1995 and 1997, were consistently significantly elevated compared with concentrations at nearby coral reefs and, at times, were significantly elevated above levels outside and in the center of the ponds. Soluble reactive phosphorus concentrations were generally low throughout the ponds but were occasionally elevated inside and at the bottom of the ponds. Mean chlorophyll a concentrations inside both ponds in 1997 were significantly elevated, indicating nutrient enrichment. Flow speeds were low at all locations in and around both mangrove ponds. Pond A (Cat Cay) has two openings, which promote flushing and exchange. Pond C (Manatee Cay) has only one opening, and therefore has greater water and nutrient retention. This retention was reflected in the significantly higher chlorophyll a concentrations inside Pond C. The two ponds are low-flow environments with natural elevation of dissolved inorganic nitrogen and sometimes dissolved phosphorus. These nutrients generate higher production inside the ponds (which support unique prop root suspension feeding communities) and influence benthic communities of these ponds and of the coral ridges at the entrances of the ponds. INTRODUCTION The Pelican Cays (16°39'N, 88°10'W) are a string of mangrove islands south of Carrie Bow Cay (CBC). Pond A (Cat Cay) and Pond C (Manatee Cay) (Fig. 1), are semi-enclosed mangrove ponds and were the focus of our descriptive study. The ponds are deep (> 10 m) circular depressions. Macintyre et al. (this volume) describe these steep-sided ponds as having 'NOAA/Marine Santuaries Division, 1305 East-West Highway, 11th Floor, SSMC-4 N/ORM62, Silver Spring, MD 20910. *Department of Biology, University of Maryland, College Park, MD 20742. 108 grown as “honeycomb shoals” on top of a drowned karst pattern that has this basic shape. The coral communities in these ponds are depauperate, but soft-bodied anthozoans such as anemones and zoanthids can be abundant. The ponds support a rich community of suspension feeders such as ascidians and sponges that often dominate the surfaces of mangrove roots along the pond edges (Goodbody, this volume; Riitzler et al., this volume). At the pond openings where water flows in and out of the ponds, shallow coral ridges (< 1 m water depth) show a gradual transition from coral dominance outside to a mixed assemblage of corals, algae, seagrasses, and soft-bodied anthozoans inside (Macintyre et al., this volume). Agaricia tenuifolia is frequently the dominant coral along this transition and can occupy large areas inside ponds as well as outside. Aronson et al. (1998) report that A. tenuifolia became the dominant coral on these reefs only in the past decade. Other corals common to the ridges are Acropora cervicornis and Porites divaricata, and the hydrocoral Millepora complanata. Mangrove communities adjacent to nearshore or island coral communities (Vijay Anand, 1995), can significantly alter local hydrodynamics of the region and are areas of high primary productivity that have the potential to locally elevate nutrient concentrations in the water column. Both dissolved inorganic nitrogen (DIN) and soluble reactive phosphorus (SRP) concentrations around a mangrove community can be significantly elevated compared with water from nearby coral reefs in the same region (Lapointe et al., 1992). Nutrients from natural or anthropogenic sources have the potential to shift benthic community structure from coral-dominated to a macroalgae and passive suspension feeding-dominated benthos. Lapointe et al. (1992) and Bell (1992) have suggested that when DIN concentrations are greater than ~1.0 uM and SRP concentrations are greater than ~0.1 «4M, benthic community structure will be dominated by macroalgae instead of corals and seagrasses. According to Bell (1992), chlorophyll a is the best indication of eutrophication and a threshold of ~0.5 g/l would be indicative for coral reef environments. With sustained nutrient elevation, macroalgae have the potential to overgrow and eliminate reef-building corals from the benthos adjacent to mangrove communities. Where corals are not overgrown by macroalgae, Atkinson et al. (1995) hypothesize, elevated nutrients could enhance coral growth. However, experiments by Marubini and Spencer-Davis (1996) have demonstrated that increased nitrate concentrations can suppress coral growth. The coral reefs at Carrie Bow Cay and in the Pelican Cays have experienced a relatively low level of anthropogenic disturbance, and the mixed reef community of coral, algae, and seagrasses at Carrie Bow Cay has been well documented (Riitzler and Macintyre, 1982). The purpose of our study was to characterize water flow, nutrient concentrations, temperature, and benthic community structure in and around two semi-enclosed mangrove ponds with such a mixed community. METHODS Study Site The study was carried out in the Pelican Cays (16° 39'N, 88° 10'W) and at Carrie Bow Cay (16° 48'N, 88° 05'W), Belize. Samples were collected from Pond A at Cat Cay and Pond C at Manatee Cay (Fig. 1). 109 Water Movement Flow speeds at locations inside and outside the ponds at Cat and Manatee Cays were measured with continuously recording Interocean S4 electromagnetic current meters. Instantaneous records are means of all 0.5-second data points, irrespective of direction. Vector- averaged records are vector-averaged means of 600 data points over a 5-min period, removing high-frequency (wave) effects on flow. Current meters were bolted to an upright PVC pipe embedded in a square cement base that could be placed in depressions among coral colonies such that the meter itself was 50 cm above the substratum. In some cases, an aluminum base was used with the same vertical PVC pipe. Water depth above the meter was 1.5 to 2.0 m. For purposes of comparison, brief deployments for several hours were carried out among habitats inside, outside, and along the tops of entrance ridges at the same time at each Cay. Long deployments lasting two days or more were also carried out on several trips during 1994-1997. Water Quality and Temperature Temperature measurements were made using Tidbit TM submersible temperature recorders (Onset Computer Inc.) programmed to record every five minutes for up to two days. Tidbits were placed at depths of 1.5 to 2.0 m depths inside and outside ponds, and at 6 m inside ponds for short periods (15—30 min), then left at inner and outer depths of 1.5—2.0 m for 48 hours. On five visits to Belize (March 1995, 1996, 1997, and July 1995, 1996) water samples were collected inside and outside both Ponds A and C for inorganic nutrient analysis. On the July 1996 and March 1996 and 1997 trips, samples were also collected at a fore-reef site near CBC for comparison with the pond sites. During each visit, we made two to four trips to Ponds A and C to collect water samples at stations throughout the ponds. The following stations were sampled on each trip to Cat and Manatee ponds: a depth of 0.5 m above the bottom in the center of the pond, 0.5 m in the center of the pond, 0.5 m above the coral ridge opening, and 0.5 m at a point 100 m outside the pond (Fig. 1). Water samples were collected in acid-washed HDPE bottles and placed on ice in the dark. Upon returning to the CBC station, we fixed a subsample for ammonium (NH4) determination with phenol solution and analyzed it within 48 hours according to methods of Parsons et al. (1984); another subsample was analyzed immediately for soluble reactive phosphorus (SRP) according to methods of Parsons et al. (1984). The remainder of the sample was filtered with a sterile GF/F filter and frozen for later analysis. The frozen subsamples were analyzed for nitrate plus nitrite (NO,+ NO,) by the Nutrient Analytical Services Laboratory at the Chesapeake Biological Laboratory in Solomons, Maryland, using standardized AutoAnalyzer methodology (D'Elia et al., 1997). Mean inorganic nutrient concentrations were calculated for each pond depth and location for each visit to Belize. Significant differences in nutrient concentrations between various locations were tested with a one-way ANOVA. Means from significant analyses were compared with an SNK test. Statistical comparisons of SRP were not performed because measures of variance could not be calculated since many measurements were below the detection limit. Chlorophyll a concentrations were measured at locations inside and outside both Ponds A and C during March 1997. Replicate samples of 4 | of seawater were collected at locations inside and outside the ponds on two occasions during March 1997. The seawater samples were filtered 110 through GF/F filters that were then placed in 30 ml of 90% acetone on ice in the dark. The filters were returned to the CBC station and extracted overnight in a refrigerator. Chlorophyll a concentrations were determined according to the method of Parsons et al. (1994). Significant differences of chlorophyll a concentrations between locations were tested with a Student's ¢-test. Benthic Community Structure Benthic community structure was characterized by videotaping 30-m transects at two locations inside and two locations outside the ponds. All transects had the coral ridge as an endpoint. A 30-m transect tape was placed along a 1- to 1.5-m depth contour at locations inside and outside the ponds. These transects were videotaped using a Yashica high 8-mm camcorder that had a scale mounted on a rod approximately 30 cm in front of the camera. A swath of substrate 30 to 50 cm wide was slowly videotaped along the 30-m transects. The videos were analyzed on a high-resolution monitor using a stop-frame, random-dot analysis (Sebens and Johnson, 1991; Aronson et al., 1994). A clear plastic sheet with 10 random dots was placed over the screen, and sessile organisms that occurred under the dots were identified and recorded; the video was then advanced to a new, nonoverlapping section for analysis. Approximately 350 dots were counted for each transect. RESULTS There was a clear temperature variation on a diel cycle for all sites except Manatee Cay, outer ridge (Fig. 2). Water exchange between the ponds and adjacent channels is slow enough that substantial solar heating occurs within the ponds, raising their temperature up to 1° C compared with greater depths in the pond (6m, Table 1) and compared with sites outside the pond (outside, Table 1). The least temperature variation was observed at the outside site at Manatee Cay, which is adjacent to a channel between cays and is thus subject to more water movement and flushing. The outside site at Cat Cay is semi-enclosed by coral ridges, although there is a very broad opening that connects this outside site with another channel between cays. Much of the water movement experienced by benthic organisms along the coral ridges and within the ponds was due to wind waves superimposed on slow unidirectional currents. Water flow was extremely low at most sites along coral ridges bordering the ponds at these cays and within the ponds themselves (Table 2). Flows on outer edges of the coral ridges at the entrances typically ranged from less than 1 cm s' to just over 3 cms on a given day (Table 2). Most sites, including the relatively unobstructed locations on the outer side of Manatee Cay, had flow speeds less than 5 cm s', but there were periods when currents could be detected, producing flows well above 5 cm s' (Fig. 3). The greatest flow speeds occur where restricted channels connect the ponds with the outer regions. Tidal or prevailing currents forcing water through those restricted channels create sustained unidirectional flows up to 12 cm s' (Fig. 4). Water flow at pond entrances can be tidally influenced, as at Pond E in Fisherman's Cay (Fig. 5), where the greatest flow speeds we observed coincided with the midtidal amplitudes and the direction of flow reversed on a tidal cycle. There was no such regular directional change at Pond C, where water flowed in and out slowly over a broad ridge, but also where wind blowing across the pond surface can have a large effect on transport. At Pond A, water flowed in through the southwestern channel during all segments of the tidal cycle, then flowed out through a large 111 Table 1. Midday temperature records at three locations each, for Manatee and Cat Cays during four days in June 1997. Outside and inside sites are as described for figure 1, inside 6 m sites are near the bottom of the ponds just below inside | m sites. Note that outside sites and inside 6 m sites were consistently slightly cooler than inside 1 m sites, but that all sites had a temperature range of one degree Celcius or less. Temperature (°C) Outside 1m Pond1m 30.1-30.4 30.7-30.9 Pond 6 m 29.9-30.1 Pond A 11:00-14:00 6/22/97 (Cat Cay) Pond C 14:00-15:00 6/22/97 30.1-30.3 30.7-30.1 30.1-30.3 (Manatee Cay) Pond A 10:00-11:00 6/24/97 29.7-29.9 30.1-30.3 29.7-29.9 (Cat Cay) Pond C 12:00-13:00 6/24/97 29.7-29.9 30.3-30.3 30.1-30.1 (Manatee Cay) Pond A 10:00-11:00 6/27/97 2993=29'5 29.3-29°5 29.3-29.5 (Cat Cay) Pond C 11:00-13:00 6/27/97 29.0-29.2 29.2-29.5 29.4-29.6 (Manatee Cay) Pond A 11:00-12:00 6/29/97 29.0-29.2 29.0-29.2 29.0-29.0 (Cat Cay) Pond C 10:00-11:00 6/29/97 29.0-29.2 29.2-29.4 29.0-29.2 (Manatee Cay) southern opening. Just after high tide, the flow speed through the southwestern opening increased, and it was at its lowest during low tide. This appeared to be a tidal flow superimposed on a unidirectional current flowing from north to south past the cays. On the inner slopes of the ponds, water flow speeds are even lower, rarely exceeding 2 cm s' (Fig. 3). Inorganic nutrient concentrations in and around Ponds A and C were variable, but dissolved inorganic nitrogen (DIN) concentrations at certain locations in the ponds were significantly elevated compared with waters from the Carrie Bow Cay fore reef, where DIN was consistently low. NO, + NO, concentrations were elevated along the coral ridge at the opening of the ponds compared with surface waters inside and outside the ponds and the Carrie Bow Cay fore-reef site (Table 3). The mean NO, + NO, concentrations along the ridges of Ponds A and C were significantly higher (p < 0.05) than those from the Carrie Bow Cay fore reef during March and July 1996 and March 1997. Mean NO,+ NO, along the ridges at the opening of Ponds A and C ranged from 0.51 to 0.71 4M while mean concentrations at Carrie Bow Cay ranged from 0.26 to 0.36 uM (Table 1). NH, concentrations were also variable in and around Ponds A and C. However, there was a pattern of NH, elevation along the ridges and at pond bottoms, whereas concentrations measured at the Carrie Bow Cay fore-reef site were consistently low. The highest concentrations of NH, were measured in the bottom of the ponds in July 1996 (Table 4). NH, concentrations were significantly higher along the pond ridge than at Carrie Bow Cay fore reef during three sampling periods of March and July 1996 and March 1997 (Table 4). Mean NH, concentrations 31.0 Cat Pond A Inside 30.5 June 22-24,1997 30.0 29.5 29.0 28.5 T a Eee 15:00 21:00 3:00 9:00, 15:00 21:00 3:00 9:00 31.0 Cat Pond A Outside 30.5 June 22-24,1997 30.0 ae a 29.5 29.0 28.5 15:00 21:00 3:00 9:00 15:00 21:00 3:00 9:00 31 2 Manatee Pond C Inside 30.5 June 27-29,1997 30.0 29.5 29.0 Temperature (°C) 28.5 1 a a aaa a aaa a 13:00 19:00 1:00 7:00 13:00 19:00 1:00 7:00 31.0 : 31 Manatee Pond C Outside 30.5 June 27-29,1997 30.0 29,5) 29.0 28.5 13:00 19:00 1:00 7:00 13:00 19:00 1:00 7:00 Time of Day Figure 2. Temperature fluctuation at four sites in the Pelican Cays during June 1997. Inner sites are at a depth of 1-2 m along the inner pond edge, and outer sites are at a similar depth outside the pond. The Cat Cay outer site is partially enclosed, whereas the Manatee Cay outer site is adjacent to an open channel between Cays, which may explain the lower diel temperature variation. along Pond A and Pond C ridges ranged from 0.20 to 0.30 uM, while the mean concentrations at Carrie Bow Cay ranged from < 0.10 to 0.15 uM. NH, in surface waters both inside and outside the ponds were generally lower than ridge and bottom concentrations. In Pond C, surface mean concentrations were significantly lower than ridge mean concentrations during the March 1996 and 1997 visits. The DIN pattern showed elevated concentrations along the entrance ridge and at times elevated concentrations at the bottom of pond. The DIN concentrations of surface water inside and outside Pond A and C were typically low and at times below detection limits. The Carrie Bow Cay fore reef had consistently low DIN concentrations (Tables 3 and 4). The SRP concentrations were generally low and did not exhibit a clear pattern at either of the ponds. During March 1996, SRP concentrations were elevated but variable in the ponds 13 Table 2. Water flow (cm sec') maxima and minima for inner and outer sites at Manatee and Cat Cays during 1995-1997. Note that flow speeds were consistently low at all sites, but that inner sites were lower than outer. Max Flow Min Flow Hours Date Site Location (cmsec') SD — (cmsec’) SD Type Deployed 3/19/95 | Manatee Cay Ridge 1.7 [QED 0.6 0.3,1.0 Vect 13 3/20/95 |Manatee Cay _— Ridge 1.5 LES eH 0.4 OLOSROM ee Wiect 24 3/21/95 |Manatee Cay _— Ridge 1.0 0.8,1.2 0.9 0.7,1.2 Vect 2 6/20/97 | Cat Cay Inner 1.6 0.0,2.4 0.9 0.4,1.6 Inst 12 6/21/97 | Cat Cay Inner Dal 0.9,3.6 0.6 0.2,1.1 Inst 24 6/22/97 | Cat Cay Inner 1.2 0.6,1.9 1.0 0.4,1.7 Inst 3 6/20/97 | Cat Cay Outer De 0.9,3.9 0.7 O:2412 Inst 12 6/21/97 | Cat Cay Outer 3.6 1.85.9 0.6 0.2,1.2 Inst 24 6/22/97 | Cat Cay Outer 2.4 0.9,4.5 0.8 0.3,0.5 Inst 3 6/27/97 | Manatee Cay Inner 2.0 L229 0.8 0.3,1.4 Inst 1] 6/28/97 | Manatee Cay Inner 2.3 NPA 30) 0.7 0.3,1.3 Inst 24 6/29/97 | Manatee Cay Inner 7s) LET E3 25 0.9 0.3,0.6 Inst 4 6/27/97 |Manatee Cay = Outer ADL) 1.7,4.4 0.8 0.3,1.4 Inst 1] 6/28/97 |Manatee Cay Outer 5,3) 3.4,7.6 0.7 0.3,1.3 Inst 24 6/29/97 |Manatee Cay Outer Dal 1.4,4.3 1.2 0.5,2.0 Inst 4 12 Pond C Inside 10 8 6 “9 4 o n oi oO my 10 ns} Oo vo mie 4) | Pond C Outside ° 10 Ge s 6 4 Otani 10:14 02:16 6/20/97 6/22/97 Figure 3. Instantaneous flow speeds at inner and outer locations at Pond A (Cat Cay). Note that the outer sites had consistently higher flow maxima. 114 Flow Speed (cm sec *') 0 4 = 25 10:15 7/9/94 Figure 4. Vector averaged flow speeds at Cat Cay coral ridge, at a point where Hise are strongest going into the pond. Note that maximum flows were generally 10-12 cms" 3-5 times higher than most other locations inside and outside the pond (Table 2). The most rapid flows at this site occurred close to the highest tide level, and were directed into the pond. 7 A 6 5 4 3 5 | 0 ‘ey co) ar E 2) S 0.7 0.75 8 0.85 0.9 0.95 1.05 om) a = 14 & fy 12 B 10 8 6 4 > 0 1.35 1.4 1.45 55) eS) 1.6 1.65 Depth (m) Figure 5. Vector averaged flow speeds. A. Fisherman’s currents flow in and out of the pond. The highest flows, 5-7 cms", at this site. B. Cat Cay western channel. The highest flows, 10-12 cm the high tide. Cay, adjacent a point where occurred at mid tide s' occurred close to Table 3. Water column NO, + NO, concentrations at Cat, Manatee and Carrie Bow Cays, Belize for five visits 3/95-3/97. Location Pond A (Cat) 3/95 Mean (SD) n Mean (SD) 7/95 3/96 Mean (SD) Mean (SD) 7/96 Mean (SD) 3/97 Bottom 0.36 (0.12) 6 0.52(0.25) 4 0.52(0.10) 3 0.51 (0.07) 4 0.44 (0.03) 3 Surface 0.38 (0.13) 6 0.51(0.08) 4 0.59(0.15) 3 0.44(0.06) 4 0.35 (0.10) 3 Ridge 0.56(0.19) 8 0.71 (0.23) 4 0.70(0.03) 6 0.62 (0.09) 4 0.52(0.11) 3 Outside 0.28 (0.09) 4 0.68(0.81) 4 0.58(0.02) 4 0.44 (0.08) 4 0.53 (0.04) 3 Pond C (Manatee) Bottom 0.37(0.15) 4 0.36(0.30) 3 0.37(0.14) 3 0.49(0.13) 4 0.23 (0.04) 3 Surface 0.13 (0.04) 4 0.33(0.08) 3 0.25(0.10) 3 0.37 (0.05) 4 0.22(0.03) 3 Ridge 0.53(0.19) 6 0.71(0.16) 5 0.51(0.21) 6 0.53(0.18) 4 0.54(0.02) 3 Outside 0.26 (0.05) 4 0.70(0.25) 6 0.43(0.06) 3 0.34 (0.10) 4 0.39(0.12) 3 Carrie Bow Barrier Reef NA NA 0.36 (0.11) 0.26 (.07) 0.29 (0.06) Table 4. Water column NH, concentrations at Cat, Manatee and Carrie Bow Cays, Belize for three visits 3/96-3/97. Location 3/96 Mean (SD) 7/96 Mean (SD) 3/97 Mean (SD) Pond A (Cat) Bottom 0.24 (0.12) 2 0.36 (0.09) 4 0.33 (0.04) 2 Surface 0.18 (0.13) 2 0.27 (0.04) 4 0.16 (0.05) I Ridge 0.20 (0.07) 4 0.27 (0.08) 4 0.24 (0.04) 2 Outside 0.17 (0.01) 4 0.21 (0.07) 4 0.15 (0.02) 2 Pond C (Manatee) Bottom 0.22 (0.03) 3} 0.32 (0.08) 4 0.16 (0.03) 2 Surface 0.12 (0.10) 2 0.20 (0.06) 4 <0.10 2 Ridge 0.20 (0.21) 4 0.31 (0.07) 4 0.29 (0.02) 2 Outside 0.15 (0.06) 2 0.17 (0.02) 4 0.13 (0.03) D Carrie Bow Cay Barrier Reef 0.11 (0.01) 0.13 (0.04) 0.13 (0.03) 3 Nits) (Table 5). High winds that occurred during this period may have resuspended nutrients from sediments in the bottom of the lagoons. At the ponds, higher concentrations of SRP were typically found inside the ponds or along the ridges (Table 5). The SRP levels from CBC were generally lower than those at all sites within the ponds for the same sampling periods. During July 1996, most mean SRP concentrations were below the detection limit (Table 5). However, individual measurements from inside the ponds were detectable. During March 1997, only the bottom water of Pond A and the bottom and surface water of Pond C had detectable concentrations of SRP. Chlorophyll a concentrations were measured inside and outside Ponds A and C. The mean 116 Table 5. Water column SRP concentrations at Cat, Manatee and Carrie Bow Cays, Belize for five visits 3/95-3/97. Location Pond A (Cat) Bottom Surface Ridge Outside Pond C (Manatee) Bottom Surface Ridge Outside Carrie Bow Barrier Reef 3/95 Mean (SD) 0.04 (0.01) 6 0.04 (0.02) 6 0.04 (0.01) 8 0.03 (0.02) 4 0.14 (0.08) 5 0.05 (0.01) 4 0.04 (0.02) 6 0.03 (0.01) 3 NA 7/95 Mean (SD) <0.03 <0).03 <0.03 <0.03 0.04 (0.01) <0).03 <0.03 <0.03 NA pH HS 7/96 3/97 3/96 Mean (SD) Mean (SD) Mean (SD) 0.07 (0.03) 2 <().03 4 0.03(0.01) 2 0.39 (0.20) 2 <0.03 4 <0.03 2 0.25(0.21) 4 <().03 4 <0.03 2 0.28 (0.26) 2 <0).03 4 <0.03 2 OLN (OS)i 2 <0.03 4 0105)(0!0)) a 0:24) (0:23), 2 <0.03 4 0.04(0.01) 2 0.44 (0.48) 6 <0.03 4 <0.03 2 <0.03 3 <0.03 4 <0.03 2 0.03 (0.01) 4 <0.03 4 <0.03 2 chlorophyll a concentrations were significantly higher (p < 0.05) inside both ponds than outside the ponds (Fig. 6) and significantly higher inside Pond C than inside Pond A. The mean (+ 1SD) chlorophyll a concentrations (g/l) at Pond A were 0.25 (+ 0.06) outside and 0.46 (+ 0.06) inside and in Pond C were 0.25 (+ 0.04) outside and 0.85 (+ 0.23) inside. 1.0 ore 0.6 aS ~) Chlorophyll a Concentration (ug/1) Pond A (Cat Cay) Outside Inside Pond C (Manatee Cay) : Outside Inside Location Figure 6. Mean (+ 1SD) concentration of chlorophyll a inside and outside Ponds A and C during July 1997 (n = 4). Inside concentrations were significantly (p <0.05) higher than outside concentrations at both locations. 117 At both Manatee and Cat Cays coral tends to be predominant on the outside of the opening to the ridge whereas macroalgae and seagrasses dominate the soft substrate inside the pond (Fig. 7). On the outside of these ponds, and along the ridge, live coral cover accounts for as much as 50% of the total area. At Manatee Cay there is a very sharp transition from coral dominance outside to macroalgae and seagrass dominance after crossing the coral ridge (Fig. 7). At Cat Cay, corals continue to persist inside Pond A but are gradually replaced by macroalgae and seagrasses (Fig. 6). A. tenuifolia is the dominant coral along all transects at Cat and Manatee Cays. Other 80 4 OLive Coral M Seagrass Macroalgae - oO > ° oO = 30 20 10 10 20 5) he ) Oy 30 30 20 10 10 20 30 Outside Ridge Inside Distance From Ridge (m) Figure 7. Percent benthic cover from combined 30 m transects along a 1-1.5 m depth contour moving from outside, onto the ridge and inside Ponds A and C at Cat and Manatee Cays, Belize, respectively (March 1995). Live corals are primarily Agaricia tenuifolia. DISCUSSION Our surveys demonstrate that Pond A in Cat Cay and Pond C in Manatee Cay are low- flow environments with some diurnal temperature fluctuations and with naturally elevated DIN concentrations in comparison with nearby reefs. Elevated DIN concentrations were measured primarily along coral ridges at the pond openings and at the bottom of the ponds. Recent investigations (Diaz and Ward, 1997; Miller-Way, personal communication) have demonstrated that common sponges and their associated cyanobacteria found on prop roots and in the benthos are capable of extremely high rates of NO, production. Other potential sources of DIN are 118 oxidation of organic materials within the reef matrix (Tribble et al., 1994), efflux from sediments (Capone et al., 1992; Williams et al., 1985), and nitrogen fixation (Wiebe et al., 1975; Capone and Carpenter, 1982), which may also contribute to the elevated DIN concentrations observed at the ponds. Concentrations of DIN and SRP were low in the central surface waters of the ponds where high chlorophyll a concentrations were measured. This indicates that phytoplankton are rapidly using DIN released into the ponds. DIN and SRP concentrations from the CBC fore reef were typically lower than levels measured in the pond. Other water quality studies in the area have also shown that DIN and SRP are barely detectable with conventional methods on fore-reef sites (Lapointe et al., 1992). Natural nutrient enrichment within the ponds is likely to influence the benthic community structure in and around the cays. The small difference in temperature between outside and inside sites at both Manatee and Cat Cays is unlikely to be an important factor explaining growth and survivorship differences of corals and other benthic organisms, except during periods of extreme high temperatures. During such periods, a difference of 1° could affect whether or not corals bleach (lose zooxanthellae and/or pigment), which could in turn affect their growth and survival. The flows inside and outside the ponds are low mainly because of the protection afforded by the mangrove islands. These relatively low flows can be extremely limiting to coral energetics (Sebens, 1997) and nutrient uptake (Thomas and Atkinson, 1997; Shyka and Lipschultz, in prep.) and may be one of the factors limiting coral distribution within the ponds. The patterns of benthic community structure, moving from outside to inside both ponds, are similar. Coral cover dominates outside the pond and onto the ridge; inside the pond, macroalgae and sea grasses are the main living components of the benthos. Lapointe et al. (1992) described a mangrove cay (Man-of-War Cay) in this region where natural nutrient elevation from a bird rookery is implicated in shifting the benthic community from a coral-dominated to macroalgae-dominated structure. The elevated concentrations of DIN, chlorophyll a, and low flow at the ponds probably play an important role in the transition in benthic community structure observed at Cat and Manatee Cay. At Cat Cay, corals were found further inside Pond A. The nutrient concentrations along the coral ridges at the two sites are not significantly different. Communities inside the ponds may differ in part because the structure of the ponds causes circulation differences. Pond C in Manatee Cay has only one opening, whereas Pond A in Cat Cay has two major openings and thus experiences increased circulation and flushing. Chlorophyll a concentrations inside Pond C are significantly higher perhaps because nutrients are retained owing to low circulation and lack of flushing. The mean chlorophyll a concentration in Pond C (0.84 g/l) is well above Bell's (1992) proposed threshold of eutrophication on coral reefs (~ 0.50 ug/l). In contrast, corals persist inside Pond A, and the mean chlorophyll a concentration was 0.45 ug/l. Our results suggest that the ponds are semi-closed systems with respect to nutrients. Other studies have demonstrated that various components of the benthic communities can release significant amounts of DIN (Diaz and Ward, 1997; Miller-Way, personal communication). The mangroves surrounding the ponds reduce flushing and exchange; hence water column production inside the ponds is higher. The high production helps support the rich community of macroalgae and suspension feeders that release DIN. Overall, nutrients and hydrodynamics play important roles in structuring the unique community found at the ponds. Elevated nutrient levels inside the ponds may not only cause a shift from coral to algal dominance and support the rich suspension-feeding communities inside the ponds, but may also have the potential to increase the growth rate of nearby corals (e.g., Atkinson et al., 1995). A low- 119 diversity but high-cover coral community was thriving along the openings of these ponds where DIN was found to be consistently elevated. However, a bleaching event in the summer of 1998 killed nearly 100% of the coral along the ridges (Precht and Aronson, personal communication). Prior to this bleaching event, A. tenuifolia dominated the substrate outside the ridges, which is also a relatively low flow environment. Reduced flow has been shown to decrease nutrient uptake (Thomas and Atkinson, 1998; Shyka and Lipschultz, in prep.) and photosynthesis and respiration of corals (Patterson et al., 1991; Lesser et al., 1994). In a separate study (Shyka et al., in prep.), we measured growth of corals in this environment. Coral growth rates in this low-flow, but nutrient-enriched, environment were very high. The mean increase in skeletal weight of Acropora cervicornis was 120% over a three-month period, March—July 1996. The mean increase in skeletal growth of A. tenuifolia was 70% over the same period. The growth rate of A. tenuifolia was significantly higher at Manatee Cay than on the fore reef at CBC at a similar depth. Manatee and Cat Cays represent unique environments in which the effects of naturally elevated nutrient concentrations may cause coral growth outside the ponds to increase, but combined with low flow and reduced circulation, may allow macroalgae and seagrass to become dominant inside the ponds. Our results suggest that the interaction of flow and nutrient concentrations plays an important role in structuring these benthic communities. CONCLUSIONS The ponds at Cat and Manatee Cays, Belize, are low-flow environments with naturally elevated water column DIN. The structure of the mangrove ponds reduces flow and promotes nutrient retention, which in turn elevates production inside the ponds. The elevated production most likely supports the unique suspension-feeding communities found in and around the ponds and promotes a macroalgae and seagrass-dominated benthic community. The elevated nutrients may also enhance growth rates of corals at the openings and outside the ponds. Overall, the natural eutrophication and restricted water movement have a strong influence on benthic community structure in and around the ponds. ACKNOWLEDGMENTS We thank Brian S. T. Helmuth, Brianna E. H. Timmerman , Deborah Danaher, Lisa Carne, Matthew Mills, Karla Heidelberg, Eugene Ariola, Jennifer Purcell, and all of the station managers at Carrie Bow Cay for assistance in the field. We are also grateful to Nutrient Analytical Services Laboratory of Chesapeake Biological Laboratory for providing nutrient analysis and equipment. We also wish to thank Matthew Mills, Brian Badgley, Assaf Gordon, and Brad Aguis for graphics and editorial assistance. This study was supported by NSF grant OCE-9302066 and a CCRE award to K. P. Sebens. This is CCR Contribution No.581. REFERENCES Aronson R. B., W. F. Precht, and I. G. Macintyre 1998. Extrinsic control of species replacement on a Holocene reef in Belize: The role of coral disease. Coral Reefs 17:223—230. 120 Atkinson M.J., B. Carlson, and G. L. Crow 1995. Coral growth in high-nutrient, low-pH seawater: A case study of corals cultured at the Waikiki Aquarium, Honolulu, Hawaii. Coral Reefs 14:215-223. Bell P2R: F: 1992. Eutophication and coral reefs (some examples in the Great Barrier Reef Lagoon. Wat Res. 26:553-S68. Capone D. G., and E. J. Carpenter 1982. Nitrogen fixation in the marine environment. Science 217:1140—-1142. Capone D. G., S. E. Dunham, S. G. Horrigan, and L. E. Duguay 1992. Microbial nitrogen transformations in unconsolidated coral reef sediments. Mar. Ecol. Prog. Ser. 80:75-88. D'Elia C.F., and W. J. Wiebe 1990. Biogeochemical nutrient cycles in coral-reef ecosystems. In Coral Reefs, edited by Z. Dubinsky, chap. 3., pp. 49-74. Amsterdam: Elsevier. D'Elia C. F., E. E. Conner, N. L. Kaumeyer, C. W. Keefe, K. V. Wood, and C. F. Zimmermann 1997. Nutrient Analytical Services Laboratory: Standard Operating Procedures. Chesapeake Biological Laboratory Technical Report Series 158-97. Diaz M. C., and B. B. Ward 1997. Sponge-mediated nitrification in tropical benthic communities. Mar. Ecol. Prog. Ser. 156:97-107. Lapointe B. E., M. M. Littler, and D. S. Littler 1992. Modification of benthic community structure by natural eutrophication: The Belize Barrier Reef. Proceedings of the 7th International Coral Reef Symposium 1:323-334. Lesser M. P., V. M. Weis, M. R. Patterson, and P. L. Jokiel 1994. Effects of morphology and water motion on carbon delivery and productivity in the reef coral, Pocillopora damicornis Linnaeus: Diffusion barriers, inorganic carbon limitation, and biochemical plasticity. J. Exp. Mar. Biol. Ecol. 178:153-179. Marubini, F., and P. S. Davies 1996. Nitrate increases zooxanthellae population density and reduces skeletogenesis in corals. Marine Biology 127:319-328. Parsons T. R., Y. Maita, and C. M. Lalli 1984. A Manual of Chemical and Biological Methods for Seawater Analysis. New York: Pergamon Press, 173 pp. Patterson M. R., R. Olson, and K. P. Sebens 1991. In situ measurements of flow effects on primary production and dark respiration in reef corals. Limnol. Oceanogr. 36:936—948 Riitzler K., and I. G. Macintyre (eds.) 1982. The Atlantic Barrier Reef at Carrie Bow Cay, Belize, 1. Structure and Communities. Washington, D.C.: Smithsonian Institution Press, 539 pp. Sebens K. P. 1997. Adaptive responses to water flow: Morphology, energetics, and distribution of reef corals. Proceedings of the 8th Coral Reef Symposium 2:1053—1058. Sebens K. P. and A. S. Johnson 1991. The effects of water movement on prey capture and distribution of reef corals. Hydrobiologia 226:91-101. Thomas F. I. M., and M. J. Atkinson 1997. Ammonium uptake by coral reefs: Effects of water velocity and surface roughness on mass transfer. Limnol. Oceanogr. 42:81-88. Tomascik T., and F. Sander 1985. Effects of eutrophication on reef-building corals. I. Growth rate of the reef-building coral Montastrea annularis. Mar. Biol. 87:143-155. Tribble G. M., M. J. Atkinson, F. J. Sansone, S. V. Smith 1994. Reef metabolism and endo-upwelling in perspective. Coral Reefs 13:199-201. Vijay Anad, P. E. 1995. Proximity of coral reefs and mangroves in the Andaman Islands. Coral Reefs 14:108. Wiebe W. J., R. E. Johannes, and K. L. Webb 1975. Nitrogen fixation in a coral reef community. Science 188:257-259. Williams S. L., S. M. Yarish, and I. P. Gill 1985b. Ammonium distributions, production, and efflux from backreef sediments, St. Croix, U.S. Virgin Islands. Mar. Ecol. Prog. Ser. 24: 57-64. ATOLL RESEARCH BULLETIN NO. 472 PHYTOPLANKTON ECOLOGY AND DISTRIBUTION AT MANATEE CAY, PELICAN CAYS, BELIZE BY STEVE L. MORTON ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 L 0 km ee5F hx - Mexico Hoye ie eS Se V7 vA, \ , / ~~ : ut NS Caribbean / N ! \ Sea i] N of Se Belize City . ~ yy oS ‘ \ ee iG Bs —— 17°N Se Ff \ Dangriga x \ x . \ SS me “3 \ \ om SES . in eat s Pond a NS = \ 5 C= Ne Des aN Guatemala Channel Bs Ped Mangrove Honduras / Ce as Soha a / Suet =i / 4 f 88° W fi / \ Sy \ Co f / / / fis I \ os 0 m 100 Figure 1. Sample sites in Pond C of Manatee Cay. PHYTOPLANKTON ECOLOGY AND DISTRIBUTION AT MANATEE CAY, PELICAN CAYS, BELIZE BY STEVE L. MORTON! ABSTRACT The phytoplankton population of Manatee Cay Pond C in the Pelican Cays of Belize differs from the population in the channel directly outside the cay in both biomass and species diversity. Significant differences also occur between the two large, semi-enclosed lobes of Pond C, C, and C,. An ecological survey demonstrated that the phytoplankton population in the pond is dominated by the mixotrophic dinoflagellate Ceratium furca. Surprisingly, diatoms common in the channel are in low abundance or absent in the Manatee pond. A large vertical migrating biomass of Gymnodinium sanguinium was found in the larger northern pond lobe known as C, INTRODUCTION The Belize Barrier Reef system is the largest in the Western Hemisphere, extending approximately 250 km from the Yucatan Peninsula to the Gulf of Honduras (Riitzler and Macintyre, 1982; James and Ginsburg, 1979). The entire barrier reef can be separated into two distinct systems. North of Belize City the shelf is shallow and has a series of islands with a discontinuous reef lacking a well-defined reef flat. South of Belize City, the reef flat is well- developed with a continuous reef. In the southern reaches of this platform the reef is cut by deep channels that form a number of shelf atolls (James and Ginsburg, 1979). The Pelican Cays group is located in this region, which is where shallow mangrove cays are immediately adjacent to channels up to 30 m deep. Several cays of this group contain central ponds separated by shallow coral ridges. These ponds, which may be 10 to 12 m deep, are characterized by a rich benthic community, primarily tunicates and sponges overgrowing mangrove substrates (Riitzler and Feller, 1996; Riitzler et al., this volume; Goodbody, this volume). Three of these cays with semi- enclosed ponds—Fisherman's Cay, Manatee Cay, and Cat Cay—are the largest examples. Because of the shallow coral ridges, the ponds are almost completely enclosed during periods of low tide. The 30-cm tide and wind-driven circulation allow only limited exchange between the semi-enclosed ponds. Villareal et al. (this volume) recently studied the complex hydrography of Manatee Cay. In this study, the water inside the pond 1s treated as separate water mass because ‘Bigelow Laboratory for Ocean Sciences, McKown Point, West Boothbay Harbor, ME 04575. Present address: National Ocean Service NOAA, 219 Ft. Johnson Rd., Charleston, SC 29412. 126 there is little or no exchange with the water outside the pond. The pond water is warmer and more saline than water outside the cay. Thus the hydrographic conditions in Manatee pond appear to respond rapidly to local heating. Villareal et al. (this volume) hypothesize that a thin layer of surface water may advect into the pond, possibly by tidal or wind-driven circulation and then become modified by local heating and evaporation to create surface water that is completely different from the adjacent water separated by the shallow coral ridge This survey reports on the phytoplankton population in two areas of the large Manatee Cay pond. This population is compared with that from the adjacent water outside Manatee Cay. MATERIALS AND METHODS During May 1996 surveys were conducted from small boats in Pond C of Manatee Cay. Samples were collected from two pond sites and one channel site (Fig. 1). The first pond site is located in the largest, northern lobe of the pond, C, (16°40.05’N 88°11.50’W). This lobe is approximately 230 m north to south and 130 m east to west. It is 35 feet deep (Urish, this volume). The second pond site is in the southern lobe of the pond, C, (16°39.95’N 88°11.51’W). This smaller lobe is approximately 120 m north to south and 100 m east to west. It has a maximum measured depth of 50 feet (Urish, this volume). The channel site outside Manatee Cay is approximately 200 m from the entrance to the ponds (16°39.93’N 88,45°11.51°W). Phytoplankton tows were made at each site using a 10-tsm-mesh net fitted with a calibrated flow meter. To reduce error caused by net clogging, only short tows of less than 2 minutes were taken. Tows were taken at three different times during the day. Samples were concentrated to a volume of 250-ml and fixed with 1% glutaraldehyde. Phytoplankton was enumerated using a 1-ml Sedwick-Rafter counting chamber. Each sample was counted three times. RESULTS Channel Site Outside Manatee Cay Seven species of centric diatoms dominated the net phytoplankton of the channel site adjacent to Manatee Cay (Table 1). The average biomass of these species was 4,512 cells * L"! with Rhizosolenia calcaravis and Bacteriastrum furcatum making up approximately 82% of the total biomass. The three species of pennate diatoms were in much lower abundance than the centric diatoms. No significant differences in species composition were noted between tows taken during the three sampling periods. Eleven species of dinoflagellates were collected from this site. The total dinoflagellate population was approximately half the diatom population, ranging from 27,700 cells * L"' to 21,800 cells * Ls The dominant type was heterotrophic dinoflagellates, Protoperidinium. No significant change in species composition or biomass was observed during the three sample times. 17) Table 1. Phytoplankton population from the channel outside the ponds of Manatee Cay Pond C (16°39.93’N 88°11.51°’W). (Each count is the mean of four separate phytoplankton tows.) Sample 1 Sample 2 Sample 3 Species (9:00 a.m.) (12:00 a.m.) (2:00 a.m.) Centric Diatom Coscinodiscus sp. 110 120 aks) Asterionellopsis glacialis 250 300 180 Rhizosolenia calcaravis 2,800 2,500 2,400 Bacteriastrum cf. furcatum 1,190 1,000 1,150 Thalassiothrix sp. 160 140 160 Thallasiosiria sp. 270 280 250 Hemialus hauckii OM) 40 64 Pennate Diatoms Pleurosigma normannii 80 75 90 Pleurosigma simonsenii 40 43 30 Navicula sp. 30 40 35 Autotrophic Dinoflagellates Ceratocorys armata 70 50 80 Dinophysis caudata 540 600 560 Prorocentrum micans 30 20 50 Gymnodinium sp. 30 20 15 Mixotrophic Dinoflagellates Ceratium tripos 430 420 400 Ceratium fusus 110 120 110 Ceratium furca 460 500 430 Ceratium trichoceros 110 120 110 Heterotrophic Dinoflagellates Protoperidinium divergens 540 550 600 Protoperidinium depressum 490 300 450 Protoperidinium pentagonum Manatee Cay Pond, Northern Lobe C, Eleven species were collected at this site (Table 2). The phytoplankton community was dominated by dinoflagellates. Ceratium furca was the dominant dinoflagellate with densities ranging from 73,790 cells * L"' to 60,000 cells * L"'. Two species, Prorocentrum hoffmannianum and Gymnodinium sanguineum, were present at this site but absent from all other sites. The centric diatoms, Cosinodiscus sp. and Rhizosolenia calcaravis, along with the cyanophyte, Oscillatoria sp., were present but in very low abundance. 128 Table 2. Phytoplankton population from Manatee Cay Pond C, northern lobe C, (16°40.05’N 88°11.50’W). (Each count is the mean of four separate phytoplankton tows.) Sample | (9:00 a.m.) Sample (12:00 a.m.) Sample 3 Species (2:00 a.m.) Diatoms Cosinodiscus sp. 15 30 9 Rhizosolenia sp. 8 5 2 Autotrophic Dinoflagellates Dinophysis caudata 670 450 550 Pyrophacus stenii 330 200 260 Prorocentrum hoffmannianum 35 20 30 Gymnodinium sanguineum 30 50 3,200 Mixotrophic Dinoflagellates Ceratium furca 73,790 64,000 60,000 Ceratium tripos 70 50 65 Heterotrophic Dinoflagellates Protoperidinium divergens 670 500 700 Protoperidinium pentagonum 80 50 20 Other Oscillatoria sp. The population of Gymnodinium sanguineum was the only species of phytoplankton that displayed a significant change during the three sampling times. Gymnodinium sanguineum increased from 30 cells * L' during the 10:00 a.m. collection to 3,200 cells * L during the 2:00 p.m. collection. Manatee Cay Pond, Southern Lobe C, Seven species of dinoflagellates were the only group of phytoplankton found at this site (Table 3). Diatoms were completely absent in all samples. Ceratium furca was the dominant species, ranging from 90,000 cells * L’! to 107,980 cells * L’'. Two species, Scrippsiella sp. and Protoceratium reticulatum, were present at this site but absent from all other sites. Diatoms were completely absent from the phytoplankton community. No significant changes in species composition and number was displayed during the three separate sampling times. 129 Table 3. Phytoplankton population from Manatee Cay Pond C, southern lobe C, (16°39.95’N, 88°11.51’W). (Each count is the mean of four separate phytoplankton tows.) Sample 3 (2:00 a.m.) Sample 1 (9:00 a.m.) Sample Species (12:00 a.m.) Autotrophic Dinoflagellates Dinophysis caudata 2,010 1,900 1,500 Scrippsiella sp. 2,610 1,600 2,840 Pyrophacus stenii 1,004 850 1,300 Mixotrophic Dinoflagellates Ceratium furca 107,980 98,000 90,000 Heterotrophic Dinoflagellates Protoperidinium divergens 1,040 Protoperidinium pentagonum 1570 DISCUSSION The phytoplankton community in the channel site is typical of a tropical oceanic community (Villareal, 1994, 1995 ; Morton, unpublished). Both the diatom and dinoflagellate species that dominate this community are commonly found throughout the Caribbean. The high dinoflagellate biomass inside the pond is remarkable, and is probably due to the unique nature of these mangrove cays. Large dinoflagellate populations have also been observed within semi- enclosed mangrove-lined lagoons north of the Pelican System. These cays include Twin Cays, Tobacco Range, and Douglas Cay (Faust and Gulledge, 1996; Morton and Villareal, in press; Morton, unpublished). At these locations, the dinoflagellate community in the mangrove-fringed lagoon is quite different from that in the adjacent water mass outside. Both the northern and southern lagoons and ponds have a dinoflagellate-dominated community without any significant diatom population. There are notable differences in the composition and numbers of dinoflagellates in the two lobes of Manatee Cay Pond C. The larger northern lobe (C,) has 11 species (8 dinoflagellate species, 2 diatom species, and 1 cyanophyte species) and a total phytoplankton population ranging from 64,856 to 75,731 cells * L''. The southern lobe (C,) has 7 species of dinoflagellates and a total phytoplankton population ranging from 98,010 to 116,214 cells * L"'. Both the northern and southern lobes are dominated by Ceratium fusus; this species contributes 92 to 97% of the total phytoplankton population. Large populations of Ceratium fusus are known to cause mortalities of benthic animals due to oxygen depletion and to cause acute toxicity of oyster larval stages (Cardwell et al., 1979; Mahoney and Steimle, 1979). These three phytoplankton populations did have a few species in common (Dinophysis caudata, Ceratium furca, Protoperidinium divergens, and P. depressum). In all cases, the two pond sites had a larger biomass than the channel site outside. The extreme example is Ceratium furca. Within the pond the population examined ranged from 60,000 to 107,980 cells * L while the outside population ranged from 100 to 120 cells * L'. 130 Three species of toxic dinoflagellates—Dinophysis caudata, Prorocentrum hoffmannianum, and Protoceratium reticulatum—were found in high numbers in both Pond C locations. Dinophysis caudata and Prorocentrum hoffmannianum produce okadaic acid and related derivatives, while Protoceratium reticulatum produces yessotoxin (Lee et al., 1987; Aikman et al., 1993; Morton et al., 1994; Satake et al., 1997). Both these lipid soluble polyether toxins have been implicated in ciguatera fish poisoning (see review by Tindall and Morton, 1998). Thus, juvenile fishes feeding inside the pond have a greater chance of accumulating toxins than fishes feeding outside. The greatest numbers of toxic dinoflagellates associated with ciguatera fish poisoning are epiphytic and not planktonic. An ecological survey of epiphytic dinoflagellates by Morton and Faust (1997) has shown a low abundance of these dinoflagellates outside Fisherman’s Cay. However, no determination of the epiphytic flora was conducted within the different ponds of the Pelican Cays. Additional sampling will be required to determine if the epiphytic dinoflagellates follow a similar trend as the planktonic dinoflagellates. The only species to display a significant variation in cell number during the three sampling periods was Gymnodinium sanguineum. This species displayed a 100-fold increase in population density between the 9:00 a.m. tow and the 2:00 p.m. tow. Gymnodinium sanguineum is a classic example of a dinoflagellate known for diurnal vertical migration and for rapid swimming. Cullen and Horrigan (1981) showed that this species can swim up to 1.1 me h'!. Villareal et al. (this volume) show that a vertically migrating chlorophyll a maximum is caused by this dinoflagellate in the northern lobe of Manatee Cay Pond C. A hydrographic survey of Manatee Cay (Villeareal et al., this volume) has shown that the water mass within the pond has a distinct temperature-salinity structure suggesting little or no exchange with the water outside the pond. This hydrographic profile would retain biomass and nutrients within the pond. The differences in the phytoplankton populations between the two sampling locations of the pond also coincide with the hydrographic survey. Each of these pond sections could be considered a different ecosystem. This survey only collected phytoplankton greater than 10 um. However, Villareal et al. (this volume) show that this fraction accounts for approximately 50 to 80% of the total chlorophyll a content. Thus, the size fraction less than 10 uum makes up a significant portion of the biomass within these ponds. Additional surveys are required to determine if the species composition of the nanoplankton and picoplankton shows similar trends as the net phytoplankton. Semi-enclosed ponds such as that of Manatee Cay appear to be prime locations for dinoflagellate blooms in traditionally nutrient-poor tropical waters. Similar Belizean mangrove-lined lagoons, Douglas Cay, Twin Cay, and Tobacco Range, also display large dinoflagellate blooms. In the red tide observed at Douglas Cay, maximum cell counts of Gonyaulax polygramma reach 3.6 x 10° cells * L' (Morton and Villareal, in press). These blooms appear to be persistent and occur independent of human activity. ACKNOWLEDGMENTS I thank Klaus Riitzler of the National Museum of Natural History, Smithsonian Institution, for his encouragement and use of the facilities at Carrie Bow Cay field station. Special thanks go to Steve Hayes and Tracy Villareal for their help sampling throughout this survey. The Smithsonian Institution Caribbean Coral Reef Ecosystem program (CCRE) supported this investigation. This paper is Contribution No. 582 of the CCRE program. 131 REFERENCES Aikman, K. A., D. R. Tindall, and S. L. Morton 1993. Physiology and potency of the dinoflagellate Prorocentrum hoffmannianum during one complete growth cycle. In Toxic Phytoplankton Blooms in the Sea, edited by T. Smayda and Y. Shimizu, 463-468. Amsterdam: Elsevier. Cardwell, R. D., S. Olsen, M. I. Carr, and E. W. Sanborn 1979. Causes of Oyster Mortality in South Puget Sound. NOAA Tech. Mem. ERL MESA- 39. Washington Department of Fisheries, Washington. Cullen, J. J., and S. G. Horrigan . 1981. Effects of nitrate on the diurnal vertical migration, carbon to nitrogen ratio, and the photosynthetic capacity of the dinoflagellate Gymnodinium splendens. Mar. Biol. 62:81-89. Ellison, A. M., E. J. Farnsworth, and R. R. Twilley 1996. Facultative mutualism between red mangroves and root-fouling sponges in Belizean mangroves. Ecol. 77:2431—2444. Faust, M. A., and R. A. Gulledge 1996. Associations of microalgae and meiofauna in floating detritus at a mangrove island, Twin Cays, Belize. J. Exp. Mar. Biol. Ecol. 197:159-175. James, N. P., and R. N. Ginsburg 1979. The seaward margin of Belize barrier and atoll reefs: Morphology, sedimentology, organism distribution and late Quaternary history. Spec. Publ. Int. Ass. Sediment. 3:i-xi, 1-191. Lee, J-S., T. Yanagi, R. Kenna, and T. Yasumoto 1987. Determination of diarrhetic shellfish toxins in various dinoflagellate species. J. Appl. Phycol. 1:147-152. Mahoney, J. B., and R. W. Steimle 1979. A mass mortality of marine animals associated with a bloom of Ceratium tripos in the New York Bight. In Toxin Dinoflagellate Blooms, edited by D. L. Taylor and H. H. Deliger, 225-230. New York: Elsevier. Morton, S. L., J. W. Bomber, and P. T. Tindall 1994. Environmental effects on the production of okadaic acid from Prorocentrum hoffmannianum Faust: I. temperature, light, and salinity. J. Exp. Mar. Biol. Ecol. 178: 67-77. Morton, S. L., and M. A. Faust 1997. Survey of toxic epiphytic dinoflagellates from the Belizean barrier reef ecosystem. Bull. Mar. Sci. 61 (3) 899-906. Morton, S. L., and T. A. Villareal In press. Bloom of Gonyaulax polygramma Stein (Dinophyceae) in a coral reef mangrove lagoon, Douglas Cay, Belize. Bull. Mar. Sci 63: 1-4. Riitzler, K., and I. C. Feller 1996. Caribbean mangrove swamps. Sci. Amer. 274:94-99. 132. Riitzler, K., and I. G. Macintyre (eds.) 1982. The Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize. I. Structure and Communities. Smithsonian Contrib. Mar. Sci. 35:1. pp. 946. Satake, M., A. L. MacKenzie, and T. Yasumoto 1997. Identification of Protoceratium reticulatum as the biogenetic origin of yessotoxin. Nat. Toxins 5:164—167. Tindall, D. R., and S. L. Morton 1998. Community dynamics and physiology of epiphytic/benthic dinoflagellate associated with ciguatera. In Physiological Ecology of Harmful Algae Blooms, edited by D. M. Anderson, A. D. Cembella, and G. M. Hallegraeff, 293-313. Berlin: Springer-Verlag. Villareal, T. A. 1994. Widespread occurrence of the Hemiaulus-cyanobacterial symbiosis in the Southwest North Atlantic Ocean. Bull. Mar. Sci. 53:1—7. 1995. Abundance and photosynthetic characteristics of Trichodesmium thiebautii along the Atlantic Barrier Reef at Carrie Bow Cay, Belize. P.S.Z.N. I. Marine Ecology 16:259-271. ATOLL RESEARCH BULLETIN NO. 473 DINOFLAGELLATE ASSOCIATIONS IN A CORAL REEF-MANGROVE ECOSYSTEM: PELICAN AND ASSOCIATED CAYS, BELIZE BY MARIA A. FAUST ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 aN f¢ 3 } 1 jf \ ene = Douglas Cay fy i} Caribbean te Sea Belize City I Dangriga { Wide 2 = 17°N j ?| Guatemala Pelican Cays 2 Elbow Cays Honduras 3 ‘ Northeast Cay 88 W Pe se Bird Cays: Ee ee} : Ridge Cay &. Manatee Cay Gb Island eiorsinyeysieisters Reef aaSseeSos Tidal flat : 15, numerous coral heads Sane SSS cs 1km Figure 1. Index map showing the location of the Pelican and associated cays DINOFLAGELLATE ASSOCIATIONS IN A CORAL REEF-MANGROVE ECOSYSTEM: PELICAN AND ASSOCIATED CAYS, BELIZE BY MARIA A. FAUST! ABSTRACT Information on the population structure of dinoflagellate taxa has been obtained from a 1994-96 study of planktonic, benthic, and oceanic dinoflagellates in the coral reef-mangrove ecosystem at Pelican Cays, Belize, and at nearby cays. Seventy-two samples were collected at six sites: Cat Cay, Douglas Cay, Elbow Cay, Fisherman’s Cay, Lagoon Cay and Manatee Cay. Up to 95% of the organisms in the >20-um microplankton samples consisted of armored dinoflagellates from 110 species in 33 genera. Of these species, 80 were photosynthetic, 30 heterotrophic, and 12 mixotrophic. Maximum cell concentrations were observed for Dinophysis caudata, Gymnodinium sanguineum, and Protoperidinium divergens. Dominant taxa included 16 Protoperidinium species, 11 Gonyaulax species, and 10 Ceratium species. Only 6 planktonic and 16 benthic species were harmful, toxin-producing dinoflagellates. Bloom-forming taxa included Ceratium furca and Gonyaulax polygramma. The findings illustrate the richness and biodiversity of dinoflagellate assemblages within the study area, as well as the importance of dinoflagellates in the microscopic food web. INTRODUCTION Dinoflagellates in coral reef-mangrove ecosystems dwell in plankton, patch reefs, seagrass beds, and sand and on the surface of macroalgae. Hence they have a complex ecology. The south central lagoon of the Belizean Barrier Reef is an oceanic coral reef-mangrove boundary environment containing a network of coral ridges and semi-enclosed or enclosed ponds, some with mangrove covers (Macintyre et al., this volume). This ecosystem, though small in geographic scale, is characterized by great typological diversity. Biological communities within this system vary markedly from one pond to another. Because of this complexity and the lack of detailed observations, few generalizations have been made about dinoflagellate distribution in Belizean coral reef-mangroves. Some important details about the associations of species in ecologically diverse environments have come to light as a result of two-week surveys of dinoflagellate species composition and abundance conducted annually from 1994 to 1996 in Pelican and nearby cays. Earlier studies have provided only limited insight into planktonic dinoflagellate associations from the southeast Caribbean Sea (Halim, 1967; Hulburt, 1968; Marshall, 1973). This discussion presents comparative information on the distribution of 'Department of Botany, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560. 136 dinoflagellates at six sites in the oceanic environment of Pelican Cays (Cat Cay, Fisherman’s Cay, Lagoon Cay, and Manatee Cay), at nearby Douglas and Elbow Cays, and at detritus-driven Twin Cays, north of Pelican Cays. METHODS The Pelican Cays, Belize, are composed of Holocene lagoon reefs (Purdy, 1994) colonized by red mangroves, Rhizophora mangle. Cut by deep channels, these reefs form a number of shelf atolls (James and Ginsburg, 1979) and an unusual network of reef ridges, both submerged and exposed (Macintyre et al., this volume). The ponds are deep in the center, have eroded peat banks round the margin, and are separated by coral ridges. Crystal clear water allows corals to proliferate adjacent to mangrove prop roots. With little water exchange from the ocean side, the ponds are warmer and more saline than usual, and could be considered separate water masses (Villareal et al., this volume). Cells were collected from six sites at the Pelican Cays and nearby: Cat Cay, Douglas Cay, Elbow Cay, Fisherman’s Cay, Lagoon Cay, and Manatee Cay (Fig.1). Study sites were selected for their varied phytoplankton associations. Seventy-two samples were gathered during yearly field trips in May between 1994 and 1996: 42 from Manatee Cay, 6 from Cat Cay, 8 from Douglas Cay, 6 from Elbow Cay, 7 from Fisherman’s Cay, and 3 from Lagoon Cay. The water temperature ranged from 26.5 to 30.6°C. Salinity levels ranged from 32 to 35.7 %o. Samples were collected just below the water surface using a 20-um pore size plankton net towed by a small boat at the lowest speed for 5 min. The net was fitted with a calibrated flow meter to estimate the volume of water entering the net. Samples were concentrated tol00-ml volume and fixed with 1% glutaraldehyde final concentration (Faust, 1990). Dinoflagellates were enumerated in a 1-ml Sedwick-Rafter counting chamber using three replicates for each sample (Guillard, 1973). Cells were identified and measured (at least 20 cells) at 630x magnification with a Carl Zeiss Axiophot light microscope. The dinoflagellate specimens generated by this study are deposited in the Dinoflagellate Collection of the U.S. National Herbarium, Smithsonian Institution. Washington, D.C. RESULTS Dinoflagellate assemblages at and near Pelican Cays included coastal benthic and oceanic species, Whereas the majority of those at Twin Cays were benthic. The majority at Pelican Cays were autotrophic and oceanic, with some neritic in origin (Table 1). Eighty species were planktonic and 30 species were benthic; both toxic and nontoxic species were present. Of the 110 dinoflagellates, 11 species are most likely new and have yet to be described. I identified 51 species from Manatee Cay, 39 from Douglas Cay, 19 from Cat Cay, 15 from Elbow Cay, 13 from Lagoon Cay. and 11 from Fisherman’s Cay. The highest numbers of species were found in samples collected from Manatee and Douglas Cays, while the lowest numbers were found in Lagoon and Fisherman's Cays (Appendix). 137 Table 1. Distribution of dinoflagellate species in the Pelican and associated Cays. | Planktonic Benthic Auto- Hetero- Taxa Amphidinium Blepharocysta Bysmatrum Ceratium Cochlodinium Coolia Corythodinium Dinophysis Diplopelta Diplosalis Diplosalopsis Gambierdiscus Goniodoma Gotoius Gonyaulax Gymnodinium Heteraulacus Lingolidinium Ostreopsis Noctiluca Peridinium Peridiniella Phaeopolykrik Plagodinium Prorocentrum Podolampas Protoceratium Protoperidiniu Pyrodinium Pyrophacus Scrippsiella Sinophysis Zygabiokonidi Total Toxic Total Toxic trophs trophs l 1 1 1 re WN We —\ — — — _ — Nn WwW — Pe (USP yy eS 14 qi ~ NN WR Re Re WR NO — a WN NWN Re bY NO Dinoflagellate assemblages in the Pelican Cays were diverse. They included coastal planktonic and benthic species and oceanic offshore species (Fig. 2). Cosmopolitan species were Ceratium furca and C. tripos, Dinophysis caudata, Gonyaulax spinifera, Prorocentrum micans, and Protoperidinium depressum. Common species were Gonyaulax spinifera and Protoperidinium sp. cf. steinni . Tropical offshore forms included Pyrodinium bahamense var. 138 bahamense and P. bahamense var. compressum; an upwelling indicator species, Ceratium symmetricum; and two rare species, Gonyaulax scrippsiae and Protoperidinium steinti (Appendix). Amphidinium F=% fAPELICAN CAYS Blepharocysta . (OTWIN CAYS Bysmatrum Ceratium Cochlodinium Coolia Corythodinium Dinophysis Diplopelta Diplosalis Diplosalopsis G ambierdiscus aes sae ee eT Se eee EEE EELS TEESE Goniodoma -=s Gotoius ssid Gonyaulax Gymnodinium Heteracaulus Lingolodinium Noctiluca Ostreopsis Peridiniella Peridinium -===scanaans pesaed Phaeopolykrikos Plagiodinium Podolampas Prorocentrum Protoceratium Protoperidinium Pyrodinium Pyrophacus Scrippsiella Sinophysis E79 ie) nN os op) & 10 12 14 16 18 NUMBER OF SPECIES Figure 2. Number of species in the dinoflagellate genera recorded in Pelican Cays and Twin Cays 139 Dinoflagellate assemblages from 33 genera were identified in the Pelican Cays ecosystem (Table 1). Dinoflagellate species are listed in the Appendix. Seventy-nine autotrophic species were from the following genera: Amphidinium, Bysmatrum, Ceratium, Cochlodinium, Coolia, Dinophysis, Goniodoma, Gambierdiscus, Gonyaulax, Noctiluca, Gymnodinium, Heteracaulus, Lingolodinium, Ostreopsis, Peridinium, Peridiniella, Phaeopolykrikos, Plagodinium, Prorocentrum, Protoceratium, Pyrodinium, Pyrophacus, and Scrippsiella. Thirty-one heterotrophic species were from the following genera: Blepharocystis, Corythodinium, Ceratium, Dinophysis, Diplosalis, Diplopelta, Diplopsalopsis, Protoperidinium, Podolampas, and Zygabiokonidium. Thirty benthic species were from the following genera: Amphidinium, Coolia, Bysmatrum, Gambierdiscus, Ostreopsis, Prorocentrum, Scrippsiella, and Sinophysis. Twelve mixotrophic species were from the following genera: Ostreopsis, Gambierdiscus, Prorocentrum, and Pyrophacus (Faust, 1998). Of the 12 mixotrophic species, 11 are known to be toxic: Ostreopsis (Norris et al., 1985), Gambierdiscus (Durant-Clement, 1987), and Prorocentrum (Murakami et al., 1982). In contrast, a total of 30 species and 14 genera were identified at Twin Cays. These included 27 autotrophic species. Twelve species were in the genus Prorocentrum and others in the following genera: Amphidinium, Bysmatrum, Ceratium, Cochlodinium, Gambierdiscus, Coolia, Dinophysis, Gonyaulax, Gymnodinium, Lingolodinium, Plagiodinium, Protoperidinium, and Scrippsiella. Two species were heterotrophic: Dinophysis rotundata and Sinophysis microcephalus (Table 1; Appendix; Faust, 1996). Few toxic species were found at Pelican Cays (Fig. 3). Of the 110 species identified, 22 are known toxin-producers: 6 of these are planktonic species and 16 are benthic species (Table 1). The number of toxic species varied in each cay. The number was higher in nutrient-enriched environments: 14 species in Manatee Cay, 12 in Douglas Cay, and 11 in Elbow Cay. Numbers were lower in oligotrophic waters: 5 species in Cat Cay, 4 in Fisherman’s Cay, and 3 in Lagoon Cay. In contrast, of the 30 species at Twin Cays, 12 were toxic (Fig. 3). Although toxic species were low at both Pelican Cays and Twin Cays, toxic populations of dinoflagellates appear to be an endemic part of Belizean mangrove ponds. The low level of toxic populations in the Southern Belizean Barrier Reef ecosystem probably prevents toxic ourbreaks of ciguatera (Yasumoto et al., 1987). The most abundant dinoflagellates were cosmopolitan species (Fig. 4). Cell numbers were an order of magnitude lower at Pelican Cays than at Twin Cays. At Manatee Cay, three autotrophic species exhibited maximum cell numbers: Ceratium furca (10,700 cells/L), Gymnodinium sanguineum (2,000 cells/L), and Dinophysis caudata (3,200 cells/L). One oceanic heterotrophic species also had elevated cell numbers: Proroperidinium divergens (1,050 cells/L). At Twin Cays, benthic autotrophic species were most abundant: Bysmatrum subsalsum (syn.=Scrippsiella subsalsum) (18,500 cells/L), Prorocentrum belizeanum (17,800 cells/L), Prorocentrum elegans (14,500 cells/L), and Prorocentrum mexicanum (10,500 cells/L). These cell populations were attached to floating detritus in protected embayments in the early afternoon on sunny and windless days. Two species, P. belizeanum (Morton et al., 1998) and P. mexicanum (Nakajima et al., 1981) have been shown to produce the toxin okadaic acid. At Pelican Cays, two autotrophic species developed "red tide" levels on three occasions, causing discoloration of the waters: at Manatee Cay, in Lagoon B, Ceratium furca reached a cell density >100,000 cells/L during May 1996 (Morton, in press); and at Douglas Cay, Gonyaulax polygramma reached concentrations of 3.5 million cells/L in May 1995, and 1.8 million cells/L in May 1996 (Morton and Villareal, in press). Both cays are nutrient enriched owing to resident pelican colonies. 140 Pelican droppings most likely create the elevated levels of organic nutrients needed to support the development of such high cell concentrations. OTOxic PELICAN CAYS faTOTAL Cat Cay Douglas Cay Elbow Cay fx Fisherman's Cay Lagoon Cay Manatee Cay TWIN CAYS 0 10 20 30 40 50 NUMBER OF SPECIES Figure 3. The total number of toxic and non-toxic dinoflagellate species at Pelican Cays and Twin Cays PELICAN CAYS C. furca D caudata B. subsalsum P belzeanum CELL CONCENTRATIONS [CELLS X 10 /L} Figure 4. Cell concentrations of most abundant dinoflagellate species at Pelican Cays and Twin Cays. 141 DISCUSSION Dinoflagellate assemblages in warm coastal waters are planktonic oceanic species (Steidinger and Williams, 1970). However, the Pelican Cays mangrove ecosystem possesses diverse assemblages that include coastal planktonic, benthic, and oceanic offshore species. Armored dinoflagellates account for 95% of these populations. Unarmored forms include only Six species in three genera: Cochlodinium, Gymnodinium, and Phaeopolykrikos (Appendix; Steidinger and Williams, 1970). The presence of oceanic species in the studied cays is an unexpected finding. Assemblages in the area inhabit ponds, which are virtually closed by coral ridges (Macintyre et al., this volume) that limit water exchange with the open ocean except during storms or extreme high tides (Villareal et al., this volume). It is difficult to ascertain whether the oceanic forms are indigenous to the Pelican Cays ecosystem or were introduced from fore-reef waters via surface currents or other means. At Twin Cays, benthic dinoflagellates are a dominant component of the dinoflagellate assemblage (Faust, 1996). They are associated with detritus (Faust and Gulledge, 1996) attached to macroalgal surfaces (Morton and Faust, 1997) and form a mucilaginous matrix (Faust, 1996). Oligotrophic waters at Pelican Cays maintain a remarkably abundant and diverse population of dinoflagellates belonging to at least 110 species. About 50% of the total species identified in this study appear to be new reports in the Belizean Barrier Reef ecosystem. However, diversity varied among the six collection sites (Table 1, Appendix, Figures 5, 6). At Elbow Cay, for example, the majority of species were autotrophic benthic, whereas at Cat Cay they were autotrophic planktonic. The highest number of species was found at Manatee Cay, which included autotrophic planktonic and benthic species, along with heterotrophic planktonic species (Table 1). Dinoflagellate diversity in Pelican Cays differs greatly from that reported in deep offshore neritic waters of the eastern Caribbean Sea (Halim, 1967; Hulburt, 1968; Marshall, 1973). The results of this study suggest that dinoflagellate abundance may be related to nutrient enrichment, as affected by the topography of each lagoon (Macintyre et al., this volume), and to the presence of an abundant in situ attached biotic component (Riitzler and Feller, 1996). The significant differences in the dinoflagellate associations at Pelican Cays and at Twin Cays (Faust, 1996) illustrate the complex ecology, species richness, biodiversity, and varied taxonomy of dinoflagellates in the coral reef-mangrove ecosystem at Belize. Oligotrophic waters at Manatee Cay exhibited three levels of dinoflagellate cell concentrations and nutrient enrichment. In the first instance, cell levels were >1,000 cells/L, a moderately enriched condition; the autotrophic species Ceratium furca, Dinophysis caudata, and Gymnodinium sanguineum, and the heterotrophic species Proroperidinium divergens were often present. In the second instance, cell levels were >100,000 cells/L, a highly enriched situation; C. furca was a bloom former (Morton, this volume). In the third instance, cell levels of many species were <1 ,000 cells/l, the most common situation. The high cell densities of C. furca were unusual. However, macroalgae, phytoplankton, invertebrates, and filter feeders were all abundant, and a constant source of dissolved nutrients (Smayda, 1991). At Twin Cays the available dissolved nutrients, originating from the decomposition of detritus, cause nutrient- enriched waters and the development of blooms (Faust, 1996). Smayda (1991) suggests that algal blooms are natural events and that elevated cell concentrations relate to local anthropogenic nutrient enrichment. In the Pelican Cay ecosystem, 142 DOUGLAS = HIDDEN LAKE 5@@ 5/92 158035 16KY¥ 199872 10KY¥ @8012 10KY 3. 00K 10. Gum Figure 5. Dinoflagellate species identified from the Pelican Cays sampling area (scanning electron micrographs; refer to the Appendix): a, Bysmatrum caponii; b, Gonyaulax grindleyi; c, Prorocentrum ruetzlerianum; d, Protoperidinium pyrum. 143 ba d 155141 10KY Figure 6. Dinoflagellate species identified from the Pelican Cays sampling area (scanning electron micrographs; refer to the Appendix): a, Protoperidinium sp. cf. steinii; b, Pyrodinium bahamense var. bahamense. dissolved organic nutrient enrichments could originate from numerous sources in addition to microalgal assemblages: the mangrove forest, corals, seagrass beds, macroalgal meadows, and peat walls dominated by filter-feeding invertebrates and macroalgae (Riitzler and Feller, 1996). In selected enclosed lagoons such as Manatee Cay, dissolved nutrients retained within the pond result in high dinoflagellate proliferation. In contrast, in semi-enclosed ponds such as those at Douglas and Elbow Cays, brown pelicans provide the organic enrichment needed for dinoflagellate blooms to develop. Here benthic dinoflagellate species G. polygramma can form populations >10° cells/L that may dominate the waters (Morton and Villareal, 1999). Nutrients generated after a "bloom" would be retained because of the very low daily tides (20-25 cm) and calm wind conditions (Ellison et al., 1996; Villareal et al., this volume). Toxic G. polygramma red tides are known to cause extensive fish and shellfish kills (Taylor, 1962). At Manatee Cay, mixotrophy was observed in an autotrophic species, Gymnodinium sanguineum, which engulfed smaller prey organisms (ciliates, pigmented nannoplankton, and microalgae). Here, G. sanguineum cells compete with heterotrophic grazers for the same food source (Bochstahler and Coats, 1993), as do harmful benthic species (Gambierdiscus, Ostreopsis and Prorocentrum) at South Water Cay (Faust, 1998). Mixotrophy is a recently described phenomenon (Jacobson and Anderson, 1986) that provides energy for cell growth, a potential advantage for dinoflagellates in nutrient-limited marine waters. As observed in this study, dinoflagellates can form "blooms" in oceanic oligotrophic mangroves. Therefore, by altering feeding behavior according to the available food sources, dinoflagellate populations can proliferate in many environments (Fenchel, 1988). The autotrophic and heterotrophic dinoflagellate population of Pelican Cays is much 144 more diverse than previously suspected. Some species are conspicuous components of plankton assemblages and some tropically categorized as benthic (Table 1). Other species are rare and present in relatively low abundance. They are important components of the plankton and may play a pivotal role in food web interactions (Fenchel, 1988). Autotrophic and heterotrophic dinoflagellates are both affected by grazers. Ciliates, nannoplankton, and flagellates have been cited as important dinoflagellate consumers (Faust and Gulledge, 1996; Jacobson and Anderson, 1986). Photosynthetic dinoflagellates consume small toxic dinoflagellates and prey on ciliates (Faust, 1998; Bockstahler and Coats, 1993; Hansen, 1991). The recent discovery of Gymnodinium sanguineum as a possible mixotroph at Manatee Cay may indicate that this species plays a dual role in the food web during high cell concentrations. The previously expected trophic role of ciliates and dinoflagellates thus appears reversed in the microscopic mangrove food web, as shown in estuarine assemblages (Bockstahler and Coats, 1993) and toxic species (Faust, 1998). At present, it is difficult to explain the observed differences between species composition and collection sites. On some occasions, blooms are so profuse that they discolor the water. These are the most noticeable instances of the association between dinoflagellates and community dynamics, but there are many less spectacular occurrences that are part of the normal seasonal succession of dinoflagellates. The latter situation may not have as obvious an impact as a toxic bloom or a red tide, but its influence on food web dynamics may still be far reaching. In view of the central role played by dinoflagellates in the coral reef-mangrove microbial food web, they are able to explicitly exploit the environment to their benefit in order to survive and proliferate. The results of this study demonstrate that the Belizean coral reef-mangrove ecosystem is a delicate and species-rich environment and, as such, should be protected and preserved. ACKNOWLEDGMENTS I thank Klaus Riitzler for his encouragement and the use of the facilities at the Smithsonian Institution’s Carrie Bow Cay field station, Belize. This investigation was supported by grants from the Smithsonian Institution, National Museum of Natural History, Caribbean Coral Reef Ecosystem program (CCRE). This paper is Contribution No. 583 of the CCRE program. REFERENCES Bockstahler, K. R., and D. W. Coats 1993: Grazing mixotrophic dinoflagellate Gymnodinium sanguineum on ciliate populations of the Chesapeake Bay. Marine Biology, 116:477-487. Ellison, A. M., E. J. Farnsworth, and R. R. Twilley 1996. Facultative mutualism between red mangroves and root-fouling sponges in Belizean mangroves. Ecology 77:2431-2444. Durant-Clement, M. 1987. Study of production and toxicity of cultured Gambierdiscus toxicus. Biological Bulletin 172:108-121. 145 Faust, M. A. 1990. Morphological details of six benthic species of Prorocentrum (Pyrrophyta) from a mangrove island, Twin Cays, Belize, including two new species. Journal of Phycology 26:548-558. 1996. Dinoflagellates in a mangrove ecosystem, Twin Cays, Belize. Nova Hedwigia 112:447-460. 1998. Mixotrophy in tropical benthic dinoflagellates. In B. Reguera, J. Blanco, L. Fernandez, and T. Wyatt (eds.), Harmful Algae, 390-394. Xunta de Galicia & Intergovernmental Oceanographic Commission of UNESCO. Faust, M. A., and R. A. Gulledge 1996. Population structure of phytoplankton and zooplankton associated with floating mangrove detritus in a mangrove island, Twin Cays, Belize. Journal of Experimental Marine Biology and Ecology 197:159-175. Fenchel, T. 1988. Marine plankton food chain. Annual Review of Ecology and Systematics 18:19-38. Guillard, R. R. L. 1973: Division rates. In Handbook of Phycological Methods. Culture Methods & Growth Measurements, edited by J. Stein, 289-311. New York: Cambridge University Press. Halim, Y. 1967. Dinoflagellates of the southeast Caribbean Sea (East-Venezuela). /nternational Revue ges. Hydrobiology 52:701-755. Hansen, P. J. 1991. Dinophysis-planktonic dinoflagellate that can act both as a prey and predator of a ciliate. Marine Ecology Progress Series 69:201-204. Hulburt, E. M. 1968. Phytoplankton observations in the Western Caribbean Sea. Bulletin of Marine Science 18:388-399. Jacobson, D., and D. M. Anderson 1986. Thecate heterotrophic dinoflagellates: feeding behavior and mechanisms. Journal of Phycology 22:249-258. James, N. P., and R. N. Ginsburg NOE, The Seaward Margin of Belize Barrier and Atoll Reefs: Morphology, Sedimentology, Organism Distribution and late Quaternary History. Special Publication of the International Association of Sediment, v. 3. Marshall, H. G. 1973. Phytoplankton observations in the Eastern Caribbean Sea. Hydrobiologia 41:45-55. Morton, S. L., and M. A. Faust LOOT: Survey of toxic epiphytic dinoflagellates from the Belizean Barrier Reef Ecosystem. Bulletin of Marine Science 61:899-906. Morton, S. L., P. D. R. Moeller, K. A. Young, and B. Lanoue 1998. Okadaic acid production from the marine dinoflagellate Prorocentrum belizeanum Faust isolated from the Belizean coral reef ecosystem. Toxicon 36:201-206. 146 Morton, S. L., and T. A. Villareal 1999: Bloom of Gonyaulax polygramma Stein (Dinophyceae) in a coral reef mangrove lagoon, Douglas Cay, Belize. Bulletin of Marine Science (in press). Murakami, Y., Y. Oshima, and T. Yasumoto 1982. Identification of okadaic acid as a toxic component of a marine flagellate Prorocentrum lima. Bulletin of Japan Society of Science and Fisheries 48:69-72. Nakajima, J., Y. Oshima, and T. Yasumoto 1981. Toxicity of benthic dinoflagellates in Okinawa. Bulletin of Japan Society of Science and Fisheries 47:1029-1033. Norris, D. R., J. W. Bomber, and E. Balech 1985. Benthic dinoflagellates associated with ciguatera from the Florida Keys. I. Ostreopsis heptagona sp. nov. In Toxic Dinoflagellates, edited by D. M. Anderson, A. W. White, and D. G. Baden, 3944. New York: Scientific. Purdy, E.G: 1994. Karst-determined facies patterns in British Honduras: Holocene carbonate sedimentation model. American Association of Petroleum Geologists 58:825-855. Ritzler, K., and I. C. Feller 1996: Caribbean mangrove swamps. Scientific American 274:94-99. Smayda, T. J. 199 ie Global epidemic of noxious phytoplankton blooms and food chain consequences in large ecosystems. In Food Chains, Yields, Models, and Management of Large Marine Ecosystems, edited by K. Sherman, L. M. Alexander, B. D., 275-307. Boulder, Colo.: Westview Press. Steidinger, K. A., and J. Williams 1970. Dinoflagellates: Memoirs of the Hourglass Cruises. Marine Research Laboratory, Florida Department of Natural Resources, St. Petersburg, Florida, v. II, 251 p. Taylor, F. J. R. 1962. Gonyaulax polygramma Stein in Cape waters: taxonomic problem related to developmental morphology. Journal of South African Botany 28:237-242. Yasumoto, Y., I. Nakajima, and R. Bagnis 1987. Finding of a dinoflagellate as a likely culprit of ciguatera. Bulletin of Japan Society of Science and Fisheries 43:1021-1026. 147 APPENDIX A list of armored dinoflagellate species recorded from the Pelican and associated Cays between 1994-1996 (* =photosynthetic species). Cat Douglas Elbow Fisherman’s Lagoon Manatee Twin Taxa Cay Cay Cay Cay Cay Cay Cays Amphidinium carterae* Blepharocystas sp.* Bysmatrum caponii* B. subsalsum* Ceratium contortum* C. furca* C. hircus* C. massilense* C. pentagonium* C. pulchellum* C. symmetricum* C. trichoceros* C. tripos* C. tripos var. atlanticum* Cochlodinium polykrikoides* Coolia monotis* +++ Hf + t+ ttt +t t+ + + + + Corythodinium sp.* Dinophysis acutoides D. caudata* D. elongatum D. mitra* D. rotundata Dinophysis sp. Diplosalis assimetrica D. bomba D. lenticula Diplopelta symmetrica Diplopelta sp. Diplopsalopsis sp. Gambierdiscus australes* G. belizeanus* G. pacificus* + + +++ + 4 G. toxicus* Goitus sp. Goniodoma sp.* Gonyaulax diacanta* G. digitalis* 148 Appendix--continued Taxa G. fragilis* G. grindleyi* G. monocanta* G. polygramma* G. reticulata* G. scrippsiae* G. spinifera* G. vrior* Gonyaulax sp.* Gymnodinium sanguineum* Gymnodinium sp. 1* Gymnodinium sp. 2* Heteraulacus sphericus Lingolidinium polyedra* Noctiluca sp.* Ostreopsis labens* O. lenticularis * O. mascarenensis* O. ovata* O. siamensis* Peridiniella spaeroidea* Peridinium venestrum* P. divergens* P. ovatum* Phaeopolykrikos sp.* Plagonidium belizeanum* Podolampas elegans Podolampas sp. Prorocentrum belizeanum* . caribbeanum* . concavum* . elegans* . emarginatum* formosum* . foraminosum* . gracile* . hoffmannianum* lima* . maculosum* . mexicanum* me ist ‘ash as) asl asl ash asl gs) ach ach a) . micans* Cat Cay Douglas Elbow Fisherman’s Cay at + + + + + Cay fe 8 de + + + + Cay Lagoon Cay Manatee Cay + + + + + + By hy EG, ay ate tet ++ + + + + Twin Cays +++ + + 4 + ++ + 4+ Appendix--continued P. norriseanum* P. ruetzlerianum* P. triestum* Prorocentrum sp.* Protoceratium excentricum* P. reticulatum Protoperidinium crassipes P. elegans P. depressum . diabolum . divergens . oblongum oceanicum pallidum pellucidum . punctulatum pyrum . quinquecorne . reticulatum 38) Fes Ae) 9) FS) AS AS Fs) AS) AS) FS . Steidingerae P. tumidum P. steinii Pyrodinium bahamense var. bahamense P. var. compressum* Pyrophacus horologium* Protoceratium sp. cf. steinii* P. steinii var. vacampoae* Scrippsiella trifida* S. trochoidea* Sinophysis microcephalus Zygabikonidium sp. Cat Cay + + + + 4+ 4+ 4+ 4 Douglas Elbow Fisherman’s Cay + + + + Cay Cay Lagoon Cay Manatee Cay + + + + + + + + + + + + + + + + ++ 4+ 4+ Twin Cays 149 ATOLL RESEARCH BULLETIN NO. 474 CHECKLIST OF MARINE ALGAE AND SEAGRASSES FROM THE PONDS OF THE PELICAN CAYS, BELIZE BY DIANE S. LITTLER, MARK M. LITTLER, AND BARRETT L. BROOKS ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 -17°00'N - e Bird Cays 7») Ridge Cay 4 bi 1.0 m). Ulvalean algal indicators of eutrophication were only observed along the eastern margin, which is otherwise unremarkable. The hanging root populations along the west side, while diverse, contain no unusual algae and are not as impressive as those of Ponds A and C (= Cat Cay Bay, Tony’s Lagoon) or Pond G (= Great Bay). A small but noteworthy pond (Pond BB = Small Pond, Fig. 1) to the south of Pond C, with a shallow entrance from the west, contains several unique features. Foremost is a spectacular population of giant Avrainvillea asarifolia (blades measure 30 cm x 20 cm; Fig. 17) at the south entrance. The feathery green Bryopsis plumosa also is exceptionally robust (Fig. 18), reaching 15 cm long here. The hanging roots contain an interesting form of the siphonaceous green Caulerpa nummularia (Fig. 20) with convex lower surfaces of the ramuli, a species that has not been found elsewhere in the Western Atlantic. Manatee Cay (Fig. 1), at the westward-facing lagoon (Pond C = Tony’s Lagoon), was surveyed in considerable detail. This is by far the most spectacular large habitat in the Pelican Cays. At the north and south of the entrance (cut off by a shallow ribbon reef of the leaf coral Agaricia tenuifolia) are uniquely large populations of the red agarophyte Gracilaria mammillaris (Fig. 10); several other commercially valuable Gracilaria spp. are present as well. A remarkable hanging population of extraordinarily large finger-like Avrainvillea digitata (Fig. 19) occurs beginning at the intertidal level of the north side of Pond C. Nearby, to the east, are gigantic brush-like Penicillus pyriformis (up to 20 cm in diameter; Fig. 21) among the Thalassia testudinum blades. The entire margin of the pond was surveyed, and all indications are of a delicate long-lived community that has undergone little human disturbance, most likely because of the shallow ribbon reef barrier across the mouth (restricting boat access) and the large volume of the pond. Interestingly, large patches of grazed T. testudinum during February 1994 and May 1995 indicated the presence of manatees. Ponds E & F (= Frenchy’s Ponds), on the southwestern region of Fisherman's Cay (Fig. 1) are rich in sessile invertebrates but tend toward algal domination. Noteworthy algal populations are draped masses of the calcareous green alga Halimeda opuntia suspended from mangrove prop roots and mound-like colonies of the paddle-shaped green alga Avrainvillea asarifolia. Many of the latter are overgrown by epiphytic Caulerpa racemosa var. occidentalis (Fig. 5). Communities on mangrove prop roots indicate little physical disturbance; however, they are not as spectacular as those of Ponds A, C, and G (= Cat Cay Bay, Tony's Lagoon, or Great Bay). Pond G (= Great Bay, Fig. 1), at the north of Fisherman’s Cay, has a north-facing entrance and is entered over a broad Thalassia testudinum flat containing isolated coral heads with scattered islets of Rhizophora mangle along the western sill. Noteworthy features are abundant 172 standing crops of commercial agarophytes, such as Gracilaria sp., G. mammillaris (Fig. 10) and the carigeenan-producer Meristiella echinocarpum (Fig. 11) along the northwestern (outer) border of the lagoon among isolated colonies of the fire coral Millepora complanata. Sheet-like Ulva rigida blades are prevalent on a nearby islet’s roots beneath a bird roosting site. Avrainvillea sp. forms a sparse aggregation on the shallow peat bank at the southern margin, beneath which are gigantic specimens of the calcareous rhizophyte Udotea cf. occidentalis (Fig. 12) extending in a strip measuring 10 m x 1.0 m. Further back among the shallow roots is an extensive patch of A. digitata. Dominant macroalgae on the roots are Acanthophora spicifera (Fig. 3), Galaxaura subverticillata (Fig. 22), and various forms of Caulerpa racemosa. Coelothrix irregularis (Fig. 2) forms dramatic neon-blue patches on submerged fallen logs. The benthic community just beneath the mangrove roots lies on a bivalve/Halimeda-hash substrate and is dominated by rhizophytic plants (Thalassia testudinum and Caulerpa racemosa covered by large mats of Ceramium sp., Caulerpa mexicana, Caulerpa sertularioides, and Acetabularia sp.; Fig. 23). Dominants on the peat bank among shallow prop roots are the blade-like Padina gymnospora (Fig. 22) and the filamentous green Caulerpa verticillata (Fig. 8). In contrast to our observations in 1992-1993, we recently found GG Ponds (= B2 Ponds), the triple-pond system of the northeast area of Fisherman's Cay (Fig. 1), to be considerably degraded, possibly by sedimentation and boat damage (physical). Colorful sponges and algae cover the mangrove prop roots of both ponds. The western margin is healthy and contains a mangrove root community dominated by algae such as shelf-like Lobophora variegata (Fig. 4). Adjacent is a seagrass bed with Bryopsidales, red and green forms of Laurencia, Coelothrix irregularis, and Ceramium sp. grading into a ribbon reef before dropping sharply to lagoon depths (24 m). Pond J (= Little Cat Bay) contains several unusually large forms of seaweeds, particularly Udotea cf. Occidentalis (Fig. 12), and has a shallow seagrass ribbon flat across its mouth. The southern portion outside of the mouth is especially rich in seaweeds that exhibit "gigantism." The mangrove prop roots on the western tip of Little Cat Cay are heavily epiphytized by the weedy red alga Acanthophora spicifera (Fig. 3). In summary, the ponds of the Pelican Cays represent spectacular, high-biodiversity, low- energy environments dominated by photosynthetic and filter-feeding populations. Most are physically delicate and highly susceptible to damage from boat wakes, physical contact (e.g., trampling), sedimentation, and nutrient enrichment. As at Twin Cays (see Taylor et al., 1986), the most delicate and palatable macroalgal forms occur on the suspended mangrove roots, a habitat that is relatively inaccessible to invertebrate grazers. Few of the ephemeral sheet-like and filamentous green algae indicative of eutrophic bird islands or anthropogenically polluted systems are present. This survey and inventory of the remarkable Pelican Cays marine plant life, where coral reef, mangrove, seagrass, and macroalgal ecosystems merge, contributes toward a baseline for conservation and management of this resource. ACKNOWLEDGMENTS We gratefully acknowledge support and funding from Harbor Branch Oceanographic Institution (HBOI Contribution No. 1332) and the Smithsonian Marine Station at Fort Pierce, Florida (SMSFP Contribution No. 489). Both organizations facilitated the laboratory aspects of this work. Funding for fieldwork was provided by NSF Grant DEBB9400534 and the Caribbean Coral Reef Ecosystem Program of the National Museum of Natural History (CCRE Contribution No. 584). 173 TAXONOMIC CHECKLIST PHYLUM RHODOPHYTA RED ALGAE ORDER: CORALLINALES FAMILY: CORALLINACEAE Amphiroa fragilissima (Linnaeus) J.V. Lamouroux 1816: 298. Basionym: Corallina fragilissima \.innaeus 1758: 806. Common: lightly attached on hard substrates, often intermixed with other species among seagrasses or in rock crevices; to 60 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30060 (US), D.&M. Littler 30061 (US), D.&M. Littler 30294 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55189 (US). Amphiroa rigida J.V. Lamouroux 1816: 297, pl. 11, fig. 1. Common: loosely attached to rock or dead coral fragments, often in seagrass beds; to 1 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55238 (US) Amphiroa sp. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55349 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30065 (US), D.&M. Littler 30271 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55209 (US). Hydrolithon farinosa f. callithamnioides (Foslie) Chamberlain 1983: 351, fig. 20b. Basionym: Melobesia farinosa f. callithamnioides Foslie 1905: 96. Common: inconspicuous; epiphytic on larger marine plants; to 15 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30280 (US). Jania adhaerens J.V. Lamouroux 1816: 270. Common. typically on hard surfaces or epiphytic on other marine plants; to 18(—35) m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30102 (US). Mesophyllum mesomorphum (Foslie) W.H. Adey 1970: 25. Basionym: Lithothamnion mesomorphum Foslie 1901: 5. Common: typically in shady cracks and crevices or epiphytic on other algae; to 35 m deep. 174 Distribution: Florida, Bahamas, Lesser Antilles, Southern Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30282 (US). ORDER: GELIDIALES FAMILY: GELIDIACEAE ~— Trichogloeopsis pedicellata (M. Howe) 1.A. Abbott & Doty 1960: 638, figs. 18-20. Basionym: Liagora pedicellata M. Howe 1920: 556. Common: typically on rocks or coral fragments, in spur-and-groove areas seaward of reef crests; to 12 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30071 (US). FAMILY: GALAXAURACEAE Galaxaura marginata (J. Ellis & Solander) J.V. Lamouroux 1816: 264. Basionym: Corallina marginata J. Ellis & Solander 1786: 115, pl. 22, fig. 6. Common. in tide pools, on shallow reef flats, or mangrove prop roots, in protected locations; to 10 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30070 (US). Galaxaura rugosa (J. Ellis & Solander) J.V. Lamouroux 1816: 263. Gametophytic Stage Basionym: Corallina rugosa J. Ellis & Solander 1786: 115, pl. 22, fig. 3. Common: on coral fragments, rocks, or mangrove prop roots, in protected areas; to 1 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30029 (US); G [Fisher- man’s Cay, Great Bay] D.&M. Littler 55216 (US). Tetrasporic Stage Basionym: Galaxaura lapidescens (J. Ellis & Solander) J.V. Lamouroux 1816: 264. Common: on coral fragments, mangrove prop roots, or rocks, in protected sandy areas; to 12 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30066 (US). Galaxaura sp. Pelican Cays Ponds: J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30221 (US), D.&M. Littler 30227 (US). Galaxaura subverticillata Kjellman 1900: 48, pl. 3, figs. 12-14; pl. 20, fig. 17. Common: typically on coral fragments or rocks, often in areas of moderate wave surge; to 10 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. 5 Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30085 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55169 (US). Tricleocarpa fragilis (Linnaeus) Huisman & R.A. Townsend 1993: 100, table 2. Basionym: Eschara fragilis Linnaeus 1758: 805. — Synonyms: Corallina oblongata J. Ellis & Solander 1786: 114, pl. 22, fig. 1; Galaxaura oblongata (J. Ellis & Solander) J.V. Lamouroux 1816: 262; Tricleocarpa oblongata (J. Ellis & Solander) Huisman & Borowitzka 1990: 168, figs. 46-49, 53-56 (see Silva et al., 1996). Common: typically on coral fragments or rocks, in protected sandy areas; to 30 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30063 (US). ORDER: GIGARTINALES FAMILY: CAULACANTHACEAE Catenella caespitosa (Withering) L.M. Irvine in Parke & Dixon 1976: 590. Basionym: Ulva caespitosa Withering 1776: 735. [Catenella repens (Lightfoot) Batters 1902: 69 (see Parke & Dixon, 1976)] Common: on mangrove prop roots, rocks, or coral fragments; extreme high intertidal. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30078 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55175 (US). Catenella sp. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55343 FAMILY: HYPNEACEAE Hypnea sp. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55073 (US), D.&M. Littler 55347 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55164 (US). FAMILY: PEYSSONNELIACEAE Peyssonnelia boergesenii Weber-van Bosse in Borgesen 1916: 137, figs. 142-145. Common: on hard substrates, often clinging to mangrove prop roots; intertidal to 40 m deep. Distribution: Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 55070 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55157 (US), D.&M. Littler 55239 (US). FAMILY: RHIZOPHYLLIDACEAE Ochtodes secundiramea (Montagne) M. Howe 1920: 583. Hypnea secundiramea Montagne 1842a: 255. Common: typically on hard substrates, in turbulent to moderately turbulent areas; to 15 m deep. Distribution: Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55227 (US). 176 FAMILY: SOLIERIACEAE Meristiella echinocarpumi (Areschoug) D.P. Cheney & P.W. Gabrielson in Gabrielson & Cheney 1987: 483, fig. 6. Basionym: Eucheuma echinocarpum Areschoug 1854: 349. Uncommon: typically on reef flats tightly adhering to substrate or as mounds in protected pristine waters; to 20 m deep. : Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30093 (US). ORDER: GRACILARIALES FAMILY: GRACILARIACEAE Gracilaria mammillaris (Montagne) M. Howe 1918: 515. Basionym: Rhodymenia mammillaris Montagne 1842a: 252. Uncommon: on rocks, mangrove prop roots, or other hard surfaces, in protected areas or exposed to moderate wave action; to 18(—60) m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55346 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30013 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55187 (US). Gracilaria sp. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55075 (US), D.&M. Littler 55353 (US), D.&M. Littler 55354 (US), D.&M. Littler 55355 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55186 (US), D.&M. Littler 55188 (US), D.&M. Littler 55204 (US). Hydropuntia cornea (J. Agardh) M.J. Wynne 1989: 476. Basionym: Gracilaria cornea J. Agardh 1852 [1851-1863]: 598. — Synonyms: Gracilaria debilis (Forsskal) Borgesen 1932: 7; Polycavernosa debilis (Forsskal) Fredericq & J.N. Norris 1985: 152 (see Wynne 1989). Common: attached to rubble fragments on protected sand-covered reef flats and turtle-grass beds; to 10 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30110 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30038 (US). Hydropuntia sp. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55345 (US). ORDER: RHODYMENIALES FAMILY: CHAMPIACEAE Champia parvula var. prostrata L.G. Williams 1951: 155. Uncommon: typically as inconspicuous epiphyte on other marine plants; to 15 m deep. Distribution: Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30259 (US). 177 FAMILY: LOMENTARIACEAE Lomentaria baileyana (Harvey) Farlow 1876: 698. Basionym: Chylocladia baileyana Harvey 1853: 185, pl. 20, fig. C. Uncommon: epiphytic on other plants; to 33 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 55063 (US); J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30211 (US). FAMILY: RHODYMENIACEAE Botryocladia shanksii E.Y. Dawson 1962: 385, pl. 1, fig. a; pl. 2, figs. a, b; pl. 5, fig. b. Uncommon: on rock or other hard surfaces in shaded habitats; intertidal to 55 m deep. Distribution: Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30072 (US). Botryocladia spinulifera W.R. Taylor & 1.A. Abbott 1973: 410, figs. 1-4. Locally abundant: inconspicuous; typically mixed in turf communities just behind reef crest on carbonate substrates; intertidal to 49 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Western Caribbean. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30115 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30295 (US). Coelothrix irregularis (Harvey) Borgesen 1920: 389, figs. 373, 374. Basionym: Cordylecladia irregularis Harvey 1853: 156. Common: forming sparse to dense mats in shaded cracks or crevices or under ledges, often on mangrove prop roots; intertidal to 10 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30111 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30057 (US). Gelidiopsis intricata (C. Agardh) Vickers 1905: 61. Basionym: Sphaerococcus intricata C. Agardh 1822 [1822-1823]: 333. Uncommon: forming large mats on mangrove prop roots or other hard surfaces; intertidal to 10 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30095 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30274 (US). Gelidiopsis planicaulis (W.R. Taylor) W.R. Taylor 1960: 353. Basionym: Wurdemannia miniata var. planicaulis W.R. Taylor 1943: 158. Common: inconspicuous; on hard surfaces, often as tufts on mangrove prop roots; to | m deep. Distribution: Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30056 (US). 178 ORDER: CERAMIALES FAMILY: CERAMIACEAE Anotrichium barbatum (J.E. Smith) Nageli 1862: 398. Basionym: Conferva barbata J.E. Smith 1807 [1790-1814]: pl. 1814. — Synonyms: Griffithsia barbata C. Agardh 1828: 132 (see Baldock 1976). Rare: inconspicuous; occurring on stones or other hard objects, often epiphytic on calcareous algae or as translucent tufts on mangrove prop roots; to 30 m deep. Distribution: Florida, Lesser Antilles, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30296 (US), D.&M. Littler 30267 (US); E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30247 (US), D.&M. Littler 30260 (US). Antithamnion sp. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30266 (US), D.&M. Littler 30037 (US); J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30216 (US). Centroceras clavulatum (C. Agardh) Montagne 1846: 140. [var. clavulatum] Common: as mats, drooping clusters, or bushy tufts on rocks, ropes, or mangrove prop roots; intertidal zone to 5 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30106 (US); B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55084 (US). Centroceras sp. Pelican Cays Ponds: E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30256 (US). Ceramium brevizonatum var. caraibicum H.E. Petersen & Borgesen in Borgesen 1924: 29, fig. 11. Common: on dead corals or epiphytic on other algae; to 1 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30018 (US). Ceramium flaccidum (Kiitzing) Ardissone 1871: 40. Basionym: Hormoceras flaccidum Kitzing 1862: 21, pl. 69, figs. a-d. — Synonyms: Ceramium byssoideum Harvey 1853: 218, nom. illeg.; C. transversale Collins & Hervey 1917: 145, pl. 5, figs. 29-31 (see Womersley 1978). Common: epiphytic on seagrasses or coarser algae; to 22 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30044 (US). Ceramium nitens (C. Agardh) J. Agardh 1851 [1851-1863]: 130. Basionym: Ceramium rubrum var. nitens C. Agardh 1824: 136. Common: on dead corals or epiphytic on other algae; to 10 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles. 179 Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55085 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30062 (US). Ceramium sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55152 (US), D.&M. Littler 55185 (US). Lejolisia exposita C.W. Schneider & Searles in Searles & Schneider 1989: 736, figs. 18-28. Uncommon: inconspicuous; epiphytic or growing as fine low turf on mangrove prop roots; lower intertidal to 32 m deep. Distribution: Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30048 (US). Spyridia filamentosa (Wulfen) Harvey in W. Hooker 1833: 337. Basionym: Fucus filamentosus Wulfen 1803: 64. Common: on solid substrates in calm protected areas; to 8 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30047 (US). Spyridia hypnoides (Bory de Saint-Vincent) Papenfuss 1968: 281. [subsp. hypnoides] Basionym: Thamnophora hypnoides Bory de Saint-Vincent 1834: 175. Uncommon: in calm protected areas; to 8 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30226 (US). Spyridia hypnoides subsp. complanata (J. Agardh) M.J. Wynne 1998: 93. Basionym: Spyridia complanata J. Agardh 1851 [1851-1863]: 343. Common: in calm protected areas, on mangrove roots, docks, and other solid structures; intertidal to 3 m deep. Distribution: Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30036 (US). Spyridia sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55190 (US), D.&M. Littler 55173 (US), D.&M. Littler 55174 (US), D.&M. Littler 55213 (US). FAMILY: DASYACEAE Dasya spinuligera Collins & Hervey 1917: 130, pl. 4, figs. 24, 25. Uncommon: on sponges or mangrove peat in protected areas; to 40 m deep. Distribution: Bahamas, Greater Antilles, Lesser Antilles, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30067 (US); E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30251. 180 Dasya spp. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55078 (US), D.&M. Littler 55090 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30045 (US), D.&M. Littler 55065 (US), D.&M. Littler 30276 (US), D.&M. Littler 30281 (US), D.&M. Littler 30298 (US). FAMILY: DELESSERIACEAE Caloglossa leprieurii (Montagne) G. Martens 1869: 234, 237. Basionym: Delesseria leprieurii Montagne 1840: 196, pl. 5, fig. 1. Common: inconspicuous; on mangrove prop roots, rocks, or other hard substrates, in protected areas; upper intertidal. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30079 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30050 (US). Hypoglossum tenuifolium (Harvey) J. Agardh 1898: 186. Basionym: Delesseria tenuifilia Harvey 1853: 97, pl. 22. Locally common: typically epiphytic or on rocks and coral fragments of deep sand plains; 45-60 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30261 (US). FAMILY: RHODOMELACEAE Acanthophora spicifera (Vahl) Borgesen 1910: 201, figs. 18, 19. Basionym: Fucus spiciferus Vahl 1802: 44. Common: early colonizer on dead coral fragments, wood substrates, pebbles, or other organisms in calm waters; intertidal to 8 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30103 (US); B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55342 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30010 (US). Bostrychia tenella (J.V. Lamouroux) J. Agardh 1863 [1851-1863]: 869. Basionym: Plocamium tenellum J.V. Lamouroux 1813: 138.— Synonyms: Fucus tenellus Vahl 1802: 45, nom. illeg.; Bostrychia binderi Harvey 1849 [1847-1849]: 68, pl. 28 [in part] (see King, Puttock & Vickery, 1988). Common: forming tightly adhering mats on rocks, seawalls, or mangrove prop roots; upper intertidal. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30059 (US), D.&M. Littler 30051 (US). 181 Bostrychia sp. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55340 (US), D.&M. Littler 55356 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55147 (US), D.&M. Littler 55217 (US), D.&M. Littler 55236 (US) Digenea simplex (Wulfen) C. Agardh 1822 [1822-1823]: 389. Basionym: Conferva simplex Wulfen 1803: 17. Common: typically on hard surfaces, often overgrown by filamentous epiphytes, abundant in heavy-surf conditions, dwarfed and denuded when buried by sand; lower intertidal to 20 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30087 (US). Herposiphonia pecten-veneris (Harvey) Falkenberg 1901: 315. Basionym: Polysiphonia pecten-veneris Harvey 1853: 46, pl. 16. Common: on hard surfaces or epiphytic on larger plants and animals; to 2 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30043 (US). Herposiphonia sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55166 (US). Laurencia gemmifera Harvey 1853: 73, pl. 18.b. Uncommon: typically on hard surfaces of shallow reef flats or attached to dead coral rubble on shallow sand plains; to 20 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55077 (US). Laurencia intricata J.V. Lamouroux 1813: 131, pl. 9, figs. 8, 9. Common: typically on rocks, shells, or coral fragments in protected sandy areas; to 3 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55079 (US); J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30223 (US). Laurencia obtusa (Hudson) J.V. Lamouroux 1813: 130. Basionym: Fucus obtusus Hudson 1778: 586. Common: typically in shallow wave-dashed habitats or areas of strong currents; to 8 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30040 (US). Laurencia papillosa (C. Agardh) Greville 1830: 52. Basionym: Chondria papillosa C. Agardh 1822 [1851-1863]: 344. Common: typically on hard surfaces exposed to moderate wave action; intertidal to 7 m deep. 182 Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55076 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30064 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55167 (US), D.&M. Littler 55201 (US). Laurencia poiteaui (J.V. Lamouroux) M. Howe 1918: 518. Basionym: Fucus poiteaui J.V. Lamouroux 1805: 63, pl. 31, figs. 2, 3. Common. typically abundant in wave-surge areas attached to rocks near base of gorgonian corals; often found in deep spur-and-groove areas on reefs; to 40 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55197 (US). Laurencia filiformis (C. Agardh) Montagne 1845: 125. Basionym: Chondria filiformis C. Agardh 1822 [1822-1823]: 358. — Synonym: Laurencia scoparia J. Agardh 1852 [1851-1863]: 746 (see Rodriguez de Rios & Saito 1982). Common: on hard substrates or mangrove prop roots; to 2 m deep. Distribution: Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30222 (US). Laurencia sp. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30074 (US); B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55344 (US); BB [Manatee Cay, Small Pond] D.&M. Littler 55047 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 55064 (US). Murrayella periclados (C. Agardh) F. Schmitz 1893: 227, footnote. Basionym: Hutchinsia periclados C. Agardh 1828: 101. Common: on mangrove prop roots, rocks, pier pilings, or seawalls in protected locations; upper intertidal. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A {Cat Cay, Cat Cay Bay] D.&M. Littler 30100 (US). Polysiphonia atlantica Kapraun & J.N. Norris 1982: 226, figs. 107a-c. Common: typically on bedrock or other hard surfaces; lower intertidal to 1 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30083 (US). Polysiphonia havanensis Montagne 1837: 352. Common: typically on hard surfaces or epiphytic on larger algae; to 1 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30213 (US). 183 Polysiphonia scopulorum Harvey 1855: 540. [var. scopulorum] Uncommon: inconspicuous; epiphytic on larger algae or seagrasses, in shallow calm waters; to 3 m deep. Distribution: Western Caribbean. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30099 (US). Polysiphonia sp. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30270 (US). PHYLUM PHAEOPHYTA BROWN ALGAE ORDER: ECTOCARPALES FAMILY: ECTOCARPACEAE Hincksia mitchelliae (Harvey) P.C. Silva in Silva et al. 1987: 73. Basionym: Ectocarpus mitchelliae Harvey 1852: 142, pl. 12, G. Common: inconspicuous; on rocks or epiphytic on other algae, often found as brown fuzz on mangrove prop roots; less than | m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30088 (US); B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55342 (US); BB [Manatee Cay, Small Pond] D.&M. Littler 55059 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30032 (US), D.&M. Littler 55059 (US), D.&M. Littler 55060 (US), D.&M. Littler 55061 (US); E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30248 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 30248 (US); J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30248 (US). ORDER: DICTYOTALES FAMILY: DICTYOTACEAE Dictyota cervicornis Kiitzing 1859: 11, pl. 24, fig. 2. Common: attached to rocks, shell fragments, or large plants in sandy shallow areas; to 3 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30107 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30054 (US). Dictyota caribaea Hornig & Schnetter in Hornig et al. 1992: 58. Synonym: Dictyota indica Kiitzing 1859: 8, pl. 17, fig. 1, sensu Vickers 1908 (see Hornig et al. 1992). Note: according to Hornig et al. (1992: 58) this entity previously was identified as Dictyota indica Kiitzing 1859: 8, pl. 17, fig. 1, sensu Vickers 1908: 39, pl. 18. However, because the type of D. indica is synonymous with D. cervicornis, a new name was assigned. Common: on rocks, other hard substrates, or mangrove peat; most commonly found in shallows around mangrove islands; to 3 m deep. 184 Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] specimen lost. Dictyota menstrualis (Hoyt) Schnetter, Hornig & Weber-Peukert 1987: 195, figs. 5, 6. Basionym: Dictyota dichotoma var. menstrualis Hoyt 1927: 616. : Common: typically on small rocks, sponges, or coral fragments in sandy areas; to 30 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30015 (US), D.&M. Littler 30269 (US). Dictyota pulchella Hornig & Schnetter 1988: 285, fig. 7. Common: on dead coral, mangrove peat, shell fragments, wood, or epiphytic on seagrasses and coarse algae in shallow areas; to 70 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30041 (US), D.&M. Littler 30055 (US); E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30258 (US); J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30225 (US). Dictyota sp. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55072 (US), D.&M. Littler 55351 (US), D.&M. Littler 55352 (US). D.&M. Littler 55357 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30017 (US), D.&M. Littler 55051 (US), D.&M. Littler 55067 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55162 (US), D.&M. Littler 55163 (US), D.&M. Littler 55165 (US), D.&M. Littler 55208 (US). D.&M. Littler 55218 (US), D.&M. Littler 55219 (US), D.&M. Littler 55220 (US), D.&M. Littler 55228 (US). Lobophora variegata (J.V. Lamouroux) Womersley ex E.C. Oliveira 1977: 217. Basionym: Dictyota variegata J.V. Lamouroux 1809a: 40. LOBOPHORA VARIEGATA HAS THREE DISTINCT FORMS DEPENDING ON DEPTH AND HABITAT CONDITIONS. DECUMBENT FORM Common: in Pelican Cays on both Peat Banks and mangrove roots, in shaded shallow areas or in deep water habitats with moderate herbivory; often dominant plant at 100 m deep; to 120 m deep. CRUST FORM Common: in Pelican Cays on mangrove roots, tightly adherent on dead coral, mangrove prop roots, or sunken logs in shallow subtidal areas where grazing is intense; to 30 m deep. RUFFLED FORM Common: in Pelican Cays lying free in Thalassia beds, in calm, shallow waters with low fish grazing; to 8 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30096 (US), D.&M. Littler 30097 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30016 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55202 (US), D.&M. Littler 55203 (US). 185 Padina payonica (Linnaeus) Thivy in W.R. Taylor 1960: 234. Basionym: Fucus pavonicus Linnaeus 1753: 1162. Common: on rocks, corals, or mangrove prop roots; found in sheltered or moderately wave- exposed areas; lower intertidal to 20 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30028 (US). Padina sanctae-crucis Borgesen 1914a: 45, figs. 27, 28 [continuous pagination: 201, figs. 153, 154]. Common: typically on rocks, shells, or dead coral on shallow reef flats; to 5 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30089 (US). Padina sp. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55074 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55215 (US); J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30248 (US). Sargassum fluitans (Borgesen) Borgesen 1914a: 66 (footnote) [continuous pagination: 222]. Basionym: Sargassum hystrix var. fluitans Borgesen 1914b: 11, fig. 8. Common: typically pelagic, floating in large clumps or rafts; major component of beach drift; one of two species characteristic of the Sargasso Sea. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30092 (US); E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30257 (US). Sargassum polyceratium var. ovatum (Collins) W.R. Taylor 1928: 129, pl. 18, figs. 7, 10; pl. 19, fig. 16. Basionym: Sargassum vulgare f. ovatum Collins 1901: 248. Common: typically on rocks in moderately turbulent habitats, often behind reef crest in rubble- pavement zone; lower intertidal to 14 m deep. Distribution: Florida, Bahamas, Greater Antilles, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30039 (US). Sargassum ramifolium Kitzing 1861: 10, pl. 32, fig. la, 1b. Uncommon: on rocks or coral fragments, often around mangrove islands; 1-3 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30091 (US). Sargassum sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55196 (US), D.&M. Littler 55199 (US), D.&M. Littler 55200 (US). Turbinaria tricostata E.S. Barton 1891: 218, pl. 54, figs. 3-4. Common: typically on rocks or dead coral fragments; in shallow areas on reef crest in strong currents or heavy wave action; lower intertidal to 1 m deep. 186 Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30090 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 55184 (US). Turbinaria turbinata (Linnaeus) Kuntze 1898: 434. Basionym: Fucus turbinatus Linnaeus 1753: 1160. Common: typically adhering tightly to hard substrates immediately behind reef crest in areas of strong turbulence; to 5 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30105 (US). PHYLUM CHLOROPHYTA GREEN ALGAE ORDER: ULVALES FAMILY: ULVACEAE Ulva rigida C. Agardh 1823 [1822-1823]: 410. Basionym: Ulva lactuca var. rigida (C. Agardh) Le Jolis 1863: 38 (see Bliding 1969). Uncommon: near bird roosting sites, typically on hard surfaces, in areas of active water motion; intertidal to 2 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: BB [Manatee Cay, Small Pond] D.&M. Littler 30262 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55170 (US). FAMILY: ULVELLACEAE Ulvella lens P. Crouan & H. Crouan 1859: 288, pl. 22, fig. E. Common: inconspicuous; on shells, hydroids, or epiphytic on other marine plants, commonly occurring on Ventricaria ventricosa; intertidal to 2 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30273 (US). ORDER: CLADOPHORALES FAMILY: ANADYOMENACEAE Anadyomene saldanhae A.B. Joly & E.C. Oliveira 1969: 30, figs. 1-3. Common: on hard substrates, sponges, or mangrove prop roots; lower intertidal to 30(—79) m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30053 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55226 (US). 187 Anadyomene sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55195 (US). Chaetomorpha sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55158 (US), D.&M. Littler 55183 (US), D.&M. Littler 55192 (US), D.&M. Littler 55225 (US). Cladophora spp. Pelican Cays Ponds: E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30254 (US), D.&M. Littler 30255 (US), D.&M. Littler 30299 (US). Rhizoclonium riparium (Roth) Kiitzing ex Harvey 1849 [1846-1851]: pl. 238. Basionym: Conferva riparia Roth 1806: 216. Common: typically on rocks, pebbles, or other hard substrates; often tangled among other species; intertidal to 1 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: J {Little Cat Cay, Little Cat Bay] D.&M. Littler 30217 (US). FAMILY: SIPHONOCLADACEAE Cladophoropsis macromeres W.R. Taylor 1928: 64, pl. 4, figs. 15, 16. Common: forming cushion-like clumps in calm shallow habitats or tangled with other algae; to 5 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30034 (US), D.&M. Littler 30268 (US). Cladophoropsis sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55211 (US). Dictyosphaeria cavernosa (Forsskal) Borgesen 1932: 2, pl. 1, fig. 1. Basionym: Ulva cavernosa Forsskal 1775: 187. Common: typically lightly attached to rocks or dead coral heads; often forming extensive mats; to 40 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30000 (US). Ventricaria ventricosa (J. Agardh) J.L. Olsen & J.A. West 1988: 104, figs. 1-4, 11. Basionym: Valonia ventricosa J. Agardh 1887: 96. Common: in cracks and crevices on hard substrates or scattered among other algae on mangrove prop roots; to 80 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30007 (US). 188 ORDER: BRYOPSIDALES FAMILY: BRYOPSIDACEAE Bryopsis hypnoides J.V. Lamouroux 1809b: 135, pl. 1, figs. 2a, 2b [also 1809a: 333]. Common: on mangrove prop roots or other hard substrates; lower intertidal to 1 m deep. Distribution: Florida, Bahamas, Greater Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30046 (US), D.&M. Littler 302745 (US). Bryopsis pennata J.V. Lamouroux 1809a: 333. [var. pennata] Common: on mangrove prop roots or other solid substrates in calm shallow waters; lower intertidal to 5 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30277 (US). Bryopsis plumosa (Hudson) C. Agardh 1823 [1822-1823]: 448. Basionym: Ulva plumosa Hudson 1778: 571. Common: typically on hard substrates, in tidepools, protected habitats, or in moderate surf behind reef crest; intertidal to 1 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30212 (US). Bryopsis ramulosa Montagne 1842b: 16, pl. 3III, fig. 2. Uncommon: inconspicuous; on mangrove prop roots or other hard surfaces; intertidal to | m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30035 (US). Bryopsis sp. Pelican Cays Ponds: BB [Manatee Cay, Small Pond] D.&M. Littler 55043 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30032 (US), D.&M. Littler 55059 (US), D.&M. Littler 55060 (US), D.&M. Littler 55061 (US); E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30248 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 30248 (US); J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30248 (US). Derbesia fastigiata W.R. Taylor 1928: 94, pl. 11, figs. 1-3. Uncommon: inconspicuous; typically epiphytic on other marine plants; intertidal to 1 m deep. Distribution: Florida, Greater Antilles, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30038 (US). Derbesia osterhoutii (L. Blinks & A. Blinks) Page 1970: 375, figs. 1-6. [Halicystis stage] Basionym: Halicystis osterhoutii L. Blinks & A. Blinks 1931: 389, pls. 22, 23, figs. 1-12, text fig. 18. Common: typically growing on crustose coralline algae such as Sporolithon or Hydrolithon, in shaded cracks and crevices; to 18 m deep. 189 Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30094 (US). FAMILY: CODIACEAE Codium decorticatum (Woodward) M. Howe 1911: 494. Basionym: Ulva decorticata Woodward 1797: 55. Common: on rock or other firm objects in protected areas; lower intertidal to 15 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55171 (US). Codium sp. Rare: typically adhering to rock or other hard surfaces, seasonally (summer) occurring off the east coast of Florida; to 20 m deep. Distribution: Florida. Pelican Cays Ponds: BB [Manatee Cay, Small Pond] D.&M. Littler 55040 (US), D.&M. Littler 55046 (US). Codium intertextum Collins & Hervey 1917: 54. Common: typically tightly adhering to rock or other solid surfaces; often forming distinct zone near low-tide mark; to 20 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30073 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55193 (US). Codium taylorii P.C. Silva 1960: 510, pl. 112, 118b, 119, 120a, 120b. Common: on mangrove prop roots, reef rubble, or other hard surfaces; to 10(—60) m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30069 (US). FAMILY: CAULERPACEAE Caulerpa cupressoides (Vahl) C. Agardh 1817: 23. [var. cupressoides] Basionym: Fucus cupressoides Vahl 1802: 38. Common: on sandy bottoms or in mangrove muds; to 3 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30021 (US). Caulerpa cupressoides var. flabellata Borgesen 1907: 368, figs. 18, 19. Uncommon: on sedimentary bottoms, anchored in fine silty sediments of mangrove lakes; to 3 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30042 (US). 190 Caulerpa macrophysa (Sonder ex Kiitzing) G. Murray 1887: 38. Basionym: Chauvinia macrophysa Sonder ex Kiitzing 1857: 6, pl. 15. Common: typically forming intertwined mats tightly attached to rock or other solid substrates, often in areas of moderate surf; intertidal to 20 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30026 (US); J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30219 (US). Caulerpa mexicana Sonder ex Kiitzing 1849: 496. Common: attached to small coral fragments or pebbles on sand or mud bottoms, in lagoons, mangroves, or seagrass beds; to 15 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: BB [Manatee Cay, Small Pond] D.&M. Littler 55144 (US), D.&M. Littler 55312 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30009 (US). Caulerpa nummularia Harvey ex J. Agardh 1873: 38. Uncommon: in low-light habitats such as shaded mangrove prop roots or under ledges in reef habitats; to 84 m deep. Distribution: Western Caribbean (Pelican Cays). Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 55071 (US). Caulerpa pusilla (Kiitzing) J. Agardh 1873: 6. Basionym: Chauvinia pusilla Kiitzing 1849: 500. Rare; inconspicuous; typically forming mats or small aggregations on deep sand plains; to 40 m deep. Distribution: Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean. Pelican Cays Ponds: J {Little Cat Cay, Little Cat Bay] D.&M. Littler 30215 (US). Caulerpa racemosa (Forsskal) J. Agardh 1873: 35. [var. racemosa] Basionym: Fucus racemosa Forsskal 1775: 191. Common: forming intertwined mats tightly adhering to rocks, in moderately heavy surf areas or in calm lagoons and bays, often present in seagrass beds; intertidal to 2 (-50) m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30088 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30032 (US), D.&M. Littler 55059 (US), D.&M. Littler 55060 (US), D.&M. Littler 55061 (US); E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30248 (US). Caulerpa racemosa var. lamourouxii (Turner) Weber-van Bosse 1898: 369, pl. 32, figs. 14. Basionym: Fucus lamourouxii Turner 1811-1819: 79, pl. 229. Uncommon: on silty substrates, generally in shallow shaded habitats such as mangrove lakes; to 30 m deep. Distribution: Bahamas, Greater Antilles, Lesser Antilles, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30033 (US), D.&M. Littler 55058 (US). 1911 Caulerpa racemosa var. occidentalis (J. Agardh) Borgesen 1907: 379, figs. 28, 29. Basionym: Caulerpa chemnitzia var. occidentalis J. Agardh 1873: 37. Common: adhering to mangrove prop roots or other hard surfaces in calm waters; to 3 m deep. Distribution: Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30026 (US), D.&M. Littler 30264 (US), D.&M. Littler 55062 (US); J [Little Cat Bay] D.&M. Littler 30220 (US). Caulerpa racemosa var. peltata (J.V. Lamouroux) Eubank in Stephenson 1944: 349. Basionym: Caulerpa peltata J.V. Lamouroux 1809a: 332. Common: on shaded mangrove prop roots, in dark crevices, or under ledges; to 5 m deep. Distribution: Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30098 (US); E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30250 (US); J [Little Cat Bay] D.&M. Littler 30214 (US). Caulerpa sertularioides (S. Gmelin) M. Howe 1905: 576. [f. sertularioides] Basionym: Fucus sertularioides S. Gmelin 1768: 151, pl. 15, fig. 4. Common: forming large stands in shallow sandy areas or on mangrove prop roots, often present in seagrass beds; to 10 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: BB [Manatee Cay, Small Pond] D.&M. Littler 55041 (US), D.&M. Littler 55143 (US), D.&M. Littler 55311 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30008 (US). Caulerpa taxifolia (H. West) C. Agardh 1817: 22. Basionym: Fucus taxifolia H. West in Vahl 1802: 36. Uncommon: growing in sand on reef flats or in fine sediments adjacent to mangrove islands in protected or moderately wave-exposed areas; to 15 m deep. Distribution: Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30104 (US); BB [Manatee Cay, Small Pond] D.&M. Littler 55042 (US); J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30218 (US). Caulerpa verticillata J. Agardh 1847: 6. [as “verticillatam’’| Common: as large aggregations on stable substrates, mangrove prop roots, or peat; often present as an understory in seagrass beds; to 30 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55358 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30022 (US). Caulerpa sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55161 (US); D.&M. Littler 55194 (US). 192 FAMILY: UDOTEACEAE Avrainvillea asarifolia Borgesen 1909: 34, fig. 4 in text, pl. 3. [f. asarifolia] Common: typically in lagoons or sand pockets between coral heads on fore-reef slopes; to 20 m deep. ; Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: BB [Manatee Cay, Small Pond] D.&M. Littler 31279 (US), E & F [Fisherman’s Cay, Frenchy’s Ponds] D.&M. Littler 30252 (US). Ayvrainvillea digitata D.S. Littler & M.M. Littler 1992: 379, fig. 3. Common: on carbonate sediments or mangrove peat, growing as large mats in shallow waters (<1 m), often interspersed among Thalassia testudinum or at the edges of mangrove islands; deeper forms (>3 m) have narrow uprights with bluntly pointed apices; Puerto Rican specimens have more club-shaped uprights; to 5 m deep. Distribution: Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30014 (US), D.&M. Littler 55054 (US); G [Fishermans Cay, Great Bay] D.&M. Littler 55235 (US). Avrainvillea longicaulis f. laxa D.S. Littler & M.M. Littler 1992: 397, fig. 13. Common: on nutrient-rich organic substrates, in interior lagoons of mangrove islands; to 2 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: BB [Manatee Cay, Small Pond] D.&M. Littler 55039 (US). Avrainvillea sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55212 (US). Halimeda discoidea Decaisne 1842: 102. Common: typically on shells, coral fragments, or other sand-covered hard surfaces; to 80 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30025 (US). Halimeda incrassata (J. Ellis) J.V. Lamouroux 1816: 307. Basionym: Corallina incrassata J. Ellis 1768: 408, pl. 17, figs. 20-27. Common: associated with seagrasses or on shallow sand flats; to 12 (-65) m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30019 (US). Halimeda monile (J. Ellis & Solander) J.V. Lamouroux 1816: 306. Basionym: Corallina monile J. Ellis & Solander 1786: 110, pl. 20, fig. c. Common: on sand flats and among seagrasses; to 30 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30114 (US). 193 Halimeda opuntia (Linnaeus) J.V. Lamouroux 1816: 308. [f. opuntia] Basionym: Corallina opuntia Linnaeus 1758: 805. Common: tightly adhering to, and forming patches on, shallow reef crests, as mounds in sand or Thalassia beds; to 25 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30001 (US), D.&M. Littler 30018 (US), D.&M. Littler 55057 (US). Halimeda simulans M. Howe 1907: 503, pl. 29. Common: often associated with mangrove peat communities or other nutrient-rich substrates; to 8 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30113 (US). Halimeda sp. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30112 (US); B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55348 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55205 (US), D.&M. Littler 55206 (US) , D.&M. Littler 55207 (US). Penicillus capitatus Lamarck 1813: 299. Common: in calm lagoons and bays on mud or sand bottoms; often intermixed with seagrasses or among mangrove prop roots; to 12 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55176 (US). Penicillus dumetosus (J.V. Lamouroux) Blainville 1834: 553. Basionym: Nesaea dumetosa J.V. Lamouroux 1816: 259, pl. 8, fig. 3a, 3b. Common: in sandy protected areas, often intermixed with seagrasses; to 15 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30109 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55177 (US), D.&M. Littler 55240 (US). Penicillus lamourouxii Decaisne 1842: 97. Common: individuals widely scattered, often intermixed with seagrasses, in calm lagoons and bays on mud or sand bottoms; to 12 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30011 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55178 (US). 194 Penicillus pyriformis A. Gepp & E. Gepp 1905: 1, pl. 468, fig. 1. [f. pyriformis] Common: on sandy bottoms in calm lagoons and bays; to 30 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30004 (US), D.&M. Littler 30006 (US). Penicillus sp. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 55052 (US). Rhipocephalus phoenix f. brevifolius A. Gepp & E. Gepp 1911: 95, pl. 31, figs. 184-186. Common: in sandy or silty areas, most commonly occurring in shallow back-reef habitats; to 20 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30003 (US), Udotea cyathiformis Decaisne 1842: 106. [f. cyathiformis] Common: in many environments from shallow mangrove peat to deep sand plains; to 30 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30002 (US), D.&M. Littler 55055 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55191 (US), D.&M. Littler 55224 (US). Udotea flabellum (J. Ellis & Solander) M. Howe 1904: 94. Basionym: Corallina flabellum J. Ellis & Solander 1786: 124, pl. 24. Common: widespread; occurring in sandy areas or seagrass beds; to 10 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30012 (US). Udotea occidentalis A. Gepp & E. Gepp 1911: 127, pl. 2, figs. 18, 22a, 22b; pl. 7, figs. 63-65. Rare: in shallow sandy areas; to 10 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30084 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 55156 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55179 (US). Udotea wilsonii A. Gepp, E. Gepp & M. Howe in Gepp & Gepp 1911: 130, 144, pl. 7, fig. 66; pl. 3, figs. 67, 68. [as ‘wilsoni’] Locally abundant: in organically rich silt or on sand plains, often growing with many thalli in close proximity due to stolonous clonal reproduction; to 18 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30005 (US). 195 ORDER: DASYCLADALES FAMILY: DASYCLADACEAE Neomeris annulata Dickie 1874: 198. Common: abundant on mangrove prop roots, coral fragments, or rocks in shallow sandy areas; intertidal to 30 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30076 (US). Neomeris sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55159 (US). FAMILY: POLYPHYSACEAE Acetabularia sp. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55160 (US). PHYLUM: CYANOPHYTA BLUE-GREEN ALGAE ORDER: OSCILLATORIALES FAMILY: OSCILLATORIACEAE Lyngbya cf. aestuarii (Martens) Liebman 1839: 492. Basionym: Conferva aesturii Martens in Jiirgens 1816: fasc. 2: no. 8. Common: as spreading mats or clumps in calm waters, appearing as wooly tufts; intertidal to 2 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55231 (US). Lyngbya cf. cladophorae Tilden 1910: 116, pl. 5, fig. 34. Uncommon: on mangrove roots or other firm objects in calm waters; intertidal to 1 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 55066 (US), D.&M. Littler 55069 (US). Lyngbya confervoides C. Agardh 1824: 73. Common: on stone or other hard surfaces, mangrove prop roots, or epiphytic on seagrasses or larger seaweeds; to 2 m deep. Distribution: Florida, Lesser Antilles, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55094 (US). 196 Lyngbya polychroa (Meneghini) Rabenhorst 1847: 83. Basionym: Leibleinia polychroa Meneghini 1844: 304. Common: epiphytic on other marine plants, often as long dark undulating masses in mangrove lagoons; intertidal to 2 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30049 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55148 (US), D.&M. Littler 55153 (US). Oscillatoria acuminata Gomont 1893: 247, pl. 7, fig. 12. Common. inconspicuous, binding surface layer of fine sediments in mangrove lakes or calm bays; to 3 m deep. Distribution: Western Caribbean. Pelican Cays Ponds: BB [Manatee Cay, Small Pond] D.&M. Littler 55050 (US). FAMILY: PHORMIDIACEAE Phormidium laysanense Lemmerman 1905: 619, pl. 7, figs. 4, 5. Uncommon: growing over other algal species or as finger-like projections from sand with basal filaments clinging to sand grains; to 2 m deep. Distribution: Greater Antilles, Lesser Antilles, Western Caribbean, Southern Caribbean. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55234 (US); J [Little Cat Cay, Little Cat Bay] D.&M. Littler 30248 (US). FAMILY: PSEUDOANABAENACEAE Spirocoleus cf. crosbyanus (Tilden) P.C. Silva in Silva et al. 1996: 62. Basionym: Phormidium crosbyanum Tilden 1909 [1894-1909]: 645. Common: forming hard button-like mounds on firm surfaces such as rocks, mangrove prop roots, or dead coral; to 2 m deep. Distribution: Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55168 (US). FAMILY: SCHIZOTHRICHACEAE Schizothrix calcicola (C. Agardh) Gomont 1890: 352. Basionym: Oscillatoria calcicola C. Agardh 1812 [1810-1812]: 37. Common: typically on rocks or other hard surfaces on shallow reef flats; to 2 m deep. Distribution: Florida, Lesser Antilles, Southern Caribbean, Western Caribbean. Pelican Cays Ponds: A [Cat Cay, Cat Cay Bay] D.&M. Littler 30084 (US); G [Fisherman’s Cay, Great Bay] D.&M. Littler 55154 (US). ORDER: STIGONEMATALES FAMILY: MASTIGOCLADACEAE Brachytrichia quoyi (C. Agardh) Bornet & Flahault 1886: 373. Basionym: Nostoc quoyi C. Agardh 1824: 22. Uncommon: epiphytic on other marine plants or on mangrove prop roots, common on pilings and breakwaters, found on rock, wood, or other firm substrates; intertidal. Distribution: Gulf of Mexico, Florida, Greater Antilles, Lesser Antilles. Pelican Cays Ponds: G [Fisherman’s Cay, Great Bay] D.&M. Littler 55145 (US). 197 PHYLUM MAGNOLIOPHYTA FLOWERING PLANTS (SEAGRASSES) ORDER: HYDROCHARITALES FAMILY: HYDROCHARITACEAE Halophila decipiens Ostenfeld 1902: 260, with fig. Common. in calm waters on soft sand or fine sedimentary bottoms; to 30 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55099 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 55053 (US). Thalassia testudinum Banks ex Konig 1805: 96. Common: abundant, conspicuous; forming extensive meadows on shallow sandy or muddy bottoms; lower intertidal to 20 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55332 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30025 (US). FAMILY: CYMODOCEACEAE Halodule wrightii Ascherson 1868: 19. Common: on sandy, soft, muddy bottoms; lower intertidal to 5 m deep. Distribution: Florida, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55333 (US); C [Manatee Cay, Tony’s Lagoon] D.&M. Littler 30024 (US). Syringodium filiforme Kiitzing in Hohenacker 1852-1962: 426. Common: widely distributed, forming dense stands in sand or fine mud sediments; to 25 m deep. Distribution: Florida, Bahamas, Greater Antilles, Lesser Antilles, Southern Caribbean, Western Caribbean, Gulf of Mexico. Pelican Cays Ponds: B [Manatee Cay, Cassiopea Cove] D.&M. Littler 55331 (US). 198 REFERENCES Abbott, I. A., and M. S. Doty 1960. Studies in the Helminthocladiaceae, II. American Journal of Botany 47(8): 632-640. Adey, W. H. 1970. A revision of the Foslie crustose coralline herbarium. Kongelige Norske Videnskabers Selskabs Skrifter 1970(1): 16. Agardh, C. A. 1810-1812. Dispositio algarum Sueciae ... Berling: Lunde [Lund]. Part 1, pp. [1]—16 (1810), part 2, pp. 17-26 (1811), part 3—S, pp. 27-45 (1812), + 3 pls. 1817. Synopsis algarum Scandinaviae ... Berling: Lunde [Lund]. xl + 135 pp. 1822-1823. Species algarum ... Vol. 1, part 2. Berling:Lundae [Lund]. [i—viii +] 169-398 (1822), 399-531 (1823) pp. 1822. Algae. Pages 1-16. In: C. S. Kunth (ed.), Synopsis plantarum, quas, in itinere ad plagam aequinoctialem orbis novi, collegerunt Al. de Humboldt et Am. Bonpland. Levrault: Paris. 1828. Species algarum ... Vol. 2. part 1. E. Martitius: Gryphiae [Greifswald]. Ixxvi+ 189 pp. Agardh, J. G. 1851-1863. Species genera et ordines algarum ... Volumen secundum: algas florideas complectens. T. O. Weigel: Lundae [Lund]. Part 1, pp. [i]—x1 + [1]-336 + 337-351 (Addenda & Index) (1851); part 2, fasc. 1, pp. 337-504 (1851); part 2, fasc. 2, pp. 505— 700 + 701-720 (Addenda & Index) (1852); part 3, fasc. 1, pp. 701-786 (1852); part 3, fasc. 2, pp. 787-1291 (1139-1158 omitted) (1863). 1873. Till algernes systematik. Nya bidrag. Lunds Universitets Ars-Skrift, Afdelningen for Mathematik och Naturvetenskap 9.71 pp. 1876. Species genera et ordines algarum ... Volumen tertium: de Florideis curae posteriores. Part 1. T. O. Weigel: Lipsiae [Leipzig]. vii + 724 pp. 1887. Till algernes systematik. Nya bidrag (Femte afdelningen). Lunds Universitets Ars- Skrift, Afdelningen for Mathematik och Naturvetenskap 23(2): 1-174, I-V pls. 1898. Species genera et ordines algarum. Voluminis tertii, pars tertia: De dispositione Delesseriearum curae posteriores. Lundae [Lund]. [vii] + 239 pp. Ardissone, F. 1871. Rivista dei Ceramii della flora italiana. Nuovo Giornale Botanico Italiano 3: 32-50. Areschoug, J. E. 1854. Phyceae novae et minus cognitae in maribus extraeuropaeis collectae quas descriptionibus atque observationibus adumbravit John Erh. Areschoug. Nova Acta Regiae Societatis Scientiarum Upsaliensis, Series 3, 1: 329-372. Ascherson, P. 1868. Mannliche bliithen von Cymodocea manatorum u. Halodule wrightii. Sitzungs-Bericht der Gesellschaft naturforschender Freunde zu Berlin. A. Effert & L. Lindtner: Berlin. 1868(16 Juni): 18-19. Baldock, R.N. 1976. The Griffithsieae group of the Ceramiaceae (Rhodophyta) and its southern Australian representatives. Australian Journal of Botany 24: 509-593. Barton, E. S. 1891. A systematic and structural account of the genus Turbinaria, Lamx. Transactions of the Linnean Society of London, Second Series, Botany 3: 215-226. 199 Batters, E. A. L. 1902. A catalogue of the British marine algae. Journal of Botany 40(suppl.): 1-107. Blainville, H. M. D. de 1834. Manuel d'actinologie ou de zoophytologie. F. G. Levrault: Paris. viii + 694 pp. Bliding, C. 1969. A critical survey of European taxa in Ulvales, II. Ulva, Ulvaria, Monostroma, Kornmannia. Botaniska Notiser 121(4): 535-629. Blinks, L. R., and A. H. Blinks 1931. Two genera of algae new to Bermuda. Bulletin of the Torrey Botanical Club 57: 389- 396. Borgesen, F. 1907. An ecological and systematic account of the Caulerpas of the Danish West Indies. Kongelige Danske Videnskabernes Selskabs Skrifter, 7. Rekke, Naturvidenskabelig og Mathematisk Afdeling 4: 337-392. 1909. The species of Avrainvilleas hitherto found on the shores of the Danish West Indies. Videnksabelige Meddelelser fra Dansk naturhistorisk Forening i Kjobenhavn 1908: 27— 44. 1910. Some new or little known West Indian Florideae. I]. Botanisk Tidsskrift 30: 177-207. 1914a. The marine algae of the Danish West Indies. Part 2. Phaeophyceae. Dansk Botanisk Arkiv 2(2): 1-66 + [2]. 1914b. The species of Sargassum found along the coasts of the Danish West Indies with remarks upon the floating forms of the Sargasso Sea. Pages 1—20. Jn: H. F. E. Jungersen and E. Warming (eds.), Mindeskrift i Anledning af Hundredaaret for Japetus Steenstrup Fodsel. Art. 32. 1916. The marine algae of the Danish West Indies. Part III. Rhodophyceae (2). Dansk Botanisk Arkiv 3: 81-144. 1920. The marine algae of the Danish West Indies. Part III. Rhodophyceae (6), with addenda to the Chlorophyceae, Phaeophyceae and Rhodophyceae. Dansk Botanisk Arkiv 3: 369- 498 + [6]. 1924. Marine algae. Jn: Plants from Beata Island, St. Domingo, collected by C. H. Ostenfeld. (Botanical Results of the Dana-Expedition 1921—22, No. 1) Dansk Botanisk Arkiv 4(7): 14-35. 1932. A revision of Forsskal's algae mentioned in Flora Aegyptiaco-Arabica and found in his herbarium in the Botanical Museum of the University of Copenhagen. Dansk Botanisk Arkiv 8(2): 1-14 + [1 index]. Bornet, E., and C. Flahault 1886. Revision des Nostocacées hétérocystées contenues dans les principaux herbiers de France. Annales des Sciences Naturelles, Botanique, Série 7, 3: 323-380. Bory de Saint-Vincent, J. B. G. M. 1834. Hydrophytes, Hydrophytae. Pages 159-178, pls. XV, XVI. In: C. Bélanger, Voyage aux Indes-Orientales, par le nord de l'Europe, les provinces du Caucase, la Géorgie, l'Arménie et la Perse, suive de détails ... pendant les années 1825, 1826, 1827, 1828 et 1829: Botanique, Ile Partie, Cryptogamie. Aurthus Bertrand: Paris. Chamberlain, Y. M. 1983. Studies in the Corallinaceae with special reference to Fosliella and Pneophyllum in the British Isles. Bulletin of the British Museum (Natural History), Botany 11(4): 291-463. Collins, F. S. 1901. The algae of Jamaica. Proceedings of the American Academy of Arts & Sciences 37: 229-270. 200 Collins, F. S., and A. B. Hervey 1917. The algae of Bermuda. Proceedings of the American Academy of Arts & Sciences 53: 1-195. Crouan, P. L., and H. M. Crouan 1859. Notice sur quelques espéces et genres nouveaux d'algues marines de la rade de Brest. Annales des Sciences Naturelles, Botanique, série 4, 12: 288-292. Decaisne, J. 1842. Essais sur une classification des algues et des polypiers calciféres de Lamouroux. Mémoire sur les corallines ou polypiers calciferes. Annales des Sciences Naturelles, Botanique, série 2, 18: 96-128. Dickie, G. 1874. On the algae of Mauritius. Journal of the Linnean Society [London], Botany 14: 190— 202. Ellis, J. 1768. Extract ofa letter from John Ellis, Esquire, F.R.S. to Dr. Linneaus, of Upsal, F.R.S. on the animal nature of the genus of zoophytes, called Corallina. Philosophical Transactions [Royal Society of London] 57: 404—420, pls. XVII, XVII. Falkenberg, P. 1901. Die Rhodomelaceen des Golfes von Neapel und der angrenzenden meerse-abschnitte. Fauna und flora des Golfes von Neapel, Monographie 26. Friedlander & Sohn: Berlin. xvi + 754 pp. Farlow, W. G. 1876. List of the marine algae of the United States. Report of the United States Fish Commissioner 1873/1875: 691-718. Foslie, M. 1901. New Melobesieae. Kongelige Norske Videnskabers Selskabs Skrifter 1901(6). 24 pp. 1905. Remarks on northern Lithothamnia. Kongelige Norske Videnskabers Selskabs Skrifter 1905(3). 138 pp. Fredericq, S., and J. N. Norris 1985. Morphological studies on some tropical species of Gracilaria Grev. (Gracilariaceae, Rhodophyta): taxonomic concepts based on reproductive morphology. Pages 137—155 In: 1. A. Abbott and J. N. Norris (eds), Taxonomy of economic seaweeds with reference to some Pacific and Caribbean species. California Sea Grant College Program [Report T-CSGCP-01 1]: La Jolla. Gabrielson, P. W., and D. P. Cheney 1987. Morphology and taxonomy of Meristiella gen. nov. (Solieriaceae, Rhodophyta). Journal of Phycology 23: 481-493. Gepp, A., and E. A. Gepp 1905. Notes on Penicillus and Rhipocephalus. Journal of Botany, British & Foreign 43: \-S. 1911. The Codiaceae of the Siboga expedition including a monograph of Flabellarieae and Udoteae. In: Siboga-expeditie Monographie 62. E. J. Brill: Leiden. 150 pp., XXII pls. Gomont, M. 1890. Essai de classification des Nostocacées homocystées. Journal de Botanique 4: 349-357. Greville, R. K. 1830. Algae britannicae ... MacLachlan & Stewart: Edinburgh. Ixxxvili + 218 pp. Grunow, A. 1867. Algae. Pages [1] + 104, pls. I, Ia, II-XI. dm: E. Fenzl (ed.), Reise der dsterreichischen Fregatte Novara um die Erde in den Jahren 1857, 1858, 1859 ... Botanischer Theil. Erster Band: Sporenpflanzen. Heft 1 K.K. Hofund Staatsdruckerei: Wien [Vienna]. 201 Harvey, W. H. 1846-1851. Phycologica britannica ... Reeve Brothers: London. Vol. 1. pp. [i-iii] [Title page (“1846”)], v—viii [Advertisement], ix—xv [List], i-vi [Index], pls. 1-120 with unpaginated text. Vol. 2. pp. [i-1i] [Title page (“1849”)], i-vi [Index], pls. 121-240 with unpaginated text. Vol. 3. pp. [i-ii] [Title page (“18517)], i1i-1v [Preface], v-xxxix [Synopsis], xli-xlv [General index], pls. 241-360 with unpaginated text. [ Plates [- LXXII (1846); LXXHI-CXLIV (1847); CXLV—CCXVI (1848); CCXVII-CCCVI (1849); CCCVII-CCCLIV (1850); CCCLV—CCCLX (1851). 1847-1849. Nereis australis. Reeve Brothers: London. Pp. i—viti + 1-64, pls. I-XXV (1847); pp. 65-124, pls. XX VI-L (1849). 1852. Nereis boreali-americana ... Part |. Melanospermeae. Smithsonian Contributions to Knowledge 3(4). ii + 150 pp., I-XII pls. 1855. Some account of the marine botany of the colony of Western Australia. Transactions of the Royal Irish Academy 22(Science): 525-566. Hohenacker, R. F. 1852-1862. Algae marinae siccatae ...Esslingen (Kirchheim from Fasc. VII). Fasc. I-XH. Nos. 1-600. [Exsiccata with printed labels.] Hooker, W. J. 1833. Class XXIV. Cryptogamia. Pages [i]—x + [1]-4 + [1]-432 pp. Jn: J. E. Smith (ed.), English flora. Vol. V. Part 1. Comprising the mosses, Hepaticae, lichens, Characeae and algae. London. Hornig, I., and R. Schnetter 1988. Notes on Dictyota dichotoma, D. menstrualis, D. indica and D. pulchella spec. nov. (Phaeophyta). Phyton 28: 277-291. HGrnig, I., R. Schnetter, W. F. Prud'homme van Reine, et al. 1992. The genus Dictyota (Phaeophyceae) in the North Atlantic. I. A new generic concept and new species. Nova Hedwigia 54: 45-62. Howe, M. A. 1904. Notes on Bahaman algae. Bulletin of the Torrey Botanical Club 31: 93-100. 1905. Phycological studies—II. New Chlorophyceae, new Rhodophyceae, and miscellaneous notes. Bulletin of the Torrey Botanical Club 32: 563-586. 1911. Phycological studies—V. Some marine algae of lower California, Mexico. Bulletin of the Torrey Botanical Club 38: 489-514. 1918. Class 3. Algae. Pages 489-540. Jn: N. L. Britton (ed.), Flora of Bermuda (illustrated). Charles Scribner's Sons: New York. 1920. Class 2. Algae. Pages 553-618. Jn: N. L. Britton and C. F. Millspaugh (eds.), The Bahama Flora. New York. Hoyt, W. D. 1927. The periodic fruiting of Dictyora and its relation to the environment. American Journal of Botany 14: 592-619. Hudson, W. 1778. Flora anglica ... Editio altera. Londini [London]. [iii +] xxxviii + 690 pp. Huisman, J. M., and R. A. Townsend 1993. An examination of Linnaean and pre-Linnaean taxa referable to Ga/axaura and Tricleocarpa (Galaxauraceae, Rhodophyta). Botanical Journal of the Linnean Society [London] 113: 95-101. Joly, A. B., and E. C. de Oliveira 1969. Notes on Brazilian algae II. A new Anadyomene of the deep water flora. Phykos 7: 27— Silk 202 Jiirgens, G. H. B. 1816 [1816-1822]. Algae aquaticae quas in littore markis Dynastium Jeveranam et Frisiam orientalem alluentis rejectas et in harum terrarum aquis habitantes collegit. Decades 1— 20. Jever. Kapraun, D. F., and J. N. Norris A 1982. The red alga Polysiphonia Greville (Rhodomelaceae) from Carrie Bow Cay and vicinity, Belize. Smithsonian Contributions to the Marine Sciences 12: 225-238. King, R. J., C. F. Puttock, and R. S. Vickery 1988. A taxonomic study on the Bostrychia tenella complex (Rhodomelaceae, Rhodophyta). Phycologia 27: 10-19. Kjellman, F. R. 1900. Om Floridé-slagtet Galaxaura, dess organografi och systematik. Konglige Svenska Vetenskaps-Akademiens Handlingar, ny foljd [series 4], No. 33. 109 pp., 20 pls. Konig, C. K. D. E. 1805. Addition to M. Cavolini's treatise on Zostera oceanica L. Kénig & Sims. Annals of Botany [London] 2: 91-98. Kuntze, O. 1898. Revisio generum plantarum ... Part 3[3]. Arthur Felix: Leipzig. i-iv + 1-576 pp. Kiitzing, F. T. 1849. Species algarum. F.A. Brockhaus: Lipsiae [Leipzig]. vi + 922 pp. 1857. Tabulae phycologicae ... Vol. 7. Nordhausen. ii + 40 pp., 100 pls. 1859. Tabulae phycologicae ... Vol. 9. Nordhausen. viii + 42 pp., 100 pls. 1861. Tabulae phycologicae ... Vol. 11. Nordhausen. [111 +] 32 pp., 100 pls. 1862. Tabulae phycologicae ... Vol. 12. Nordhausen. iv + 30 pp., 100 pls. Lamarck, J. B. de 1813. Sur les polypiers empatés. Annales du Muséum d'Historie Naturelle [Paris] 20(7): 294— B12: Lamouroux, J. V. F. 1805. Dissertations sur plusieurs espéeces de Fucus ... Agen. xxiv + 83 [85 errata] pp., XXXVI pls. 1809a. Exposition des caractéres du genre Dictyota, et tableau des especes qu'il renferme. Journal de Botanique [Desvaux] 2: 38-44. 1809b. Mémoire sur trois nouveaux generes de la famille des algues marines. Journal de Botanique [Desvaux] 2: 129-135, pl. 1; pl. HI: fig. 1. 1813. Essai sur les genres de la famille des thallassiophytes non articulées. Annales du Muséum d'Historie Naturelle [Paris] 20: 21-47, 115-139, 267-293. Lapointe, B. E., M. M. Littler, and D. S. Littler 1993. Modification of benthic community structure by natural eutrophication: the Belize barrier reef. Proceedings of the seventh International Coral Reef Symposium, Guam, 19925 NW 323-334) Lemmermann, E. 1905. Die Algenflora der Sandwich-Inseln. Ergebnisse einer Reise nach dem Pacific. H. Schauinsland 1896/97. Botanische Jahrbiicher fiir Systematik, Pflanzengeschichte und Pflanzengeographie 34: 607-663, pls. VII, VIII. Linnaeus, C. 1753. Species plantarum ... Laurentii Salvii: Holmia [Stockholm]. 561—1200 + [31] pp. 1758. Systema naturae per regna tria naturae ... Editio decima ... Vol. 1. Salvius: Holmiae [Stockholm]. iv + 823 pp. 203 Littler, D. S., and M. M. Littler 1990. Systematics of Udotea species (Bryopsidales, Chlorophyta) in the tropical western Atlantic. Phycologia 29: 206-252. 1992. Systematics of Avrainvillea (Bryopsidales, Chlorophyta) in the tropical western Atlantic. Phycologia 31: 375-418. 1997. An illustrated marine flora of the Pelican Cays, Belize. Bulletin of the Biological Society of Washington 9: 1-149. Littler, D. S., M. M. Littler, and B. L. Brooks 1995. Marine algae and seagrasses from the Tobacco Range Fracture Zone, Belize, C.A. Afoll Research Bulletin, No. 429. 43 pp. Littler, D. S., M. M. Littler, K. E. Bucher, and J. N. Norris 1989. Marine plants of the Caribbean, a field guide from Florida to Brazil. Smithsonian Institution Press: Washington, D.C. 272 pp. Littler, M. M., D. S. Littler, and B. E. Lapointe 1993. Modification of tropical reef community structure due to cultural eutrophication: the southwest coast of Martinique. Proceedings of the Seventh International Coral Reef Symposium, Guam, 1992. 1: 335-343. Littler, M. M., P. R. Taylor, D. S. Littler, R. H. Sims, and J. N. Norris 1985. The distribution, abundance and primary productivity of submerged macrophytes in the Belize barrier reef mangrove system. Afoll Research Bulletin, No. 289, 1-20. Littler, M. M., P. R. Taylor, D. S. Littler, R. H. Sims, and J. N. Norris 1987. Dominant macrophyte standing stocks, productivity and community structure on a Belizean barrier-reef. Atoll Research Bulletin, No. 302, 1-24. Martens, G. von 1869. Beitrage zur Algen-Flora Indiens. Flora 52: 233-238. Meneghini, G. 1844. Algarum species novae vel minus notae a prof. J. Meneghini propositae. Giornale Botanico Italiano, Anno 1, 1: 296-306. Montagne, C. 1837. Centurie de plantes cellulaires exotiques nouvelles. Annales des Sciences Naturelles, Botanique, série 2, 8: 345-370. 1840. Seconde centurie de plantes cellulaires exotiques nouvelles. Décades I et II. Annales des Sciences Naturelles, Botanique, série 2, 13: 193-207, pl. 5; pl. 6: figs. 1, 3. 1842a. Troisiéme centurie de plantes cellulaires exotiques nouvelles. Décades V, VI, VII et VIII. Annales des Sciences Naturelles, Botanique, série 2, 18: 241-256. 1842b. Botanique. Plantes cellulaires. Pages [v] + x + 549, Atlas: pls. I-XX. Jn: R. de la Sagra (ed.), Histoire physique, politique et naturelle de l'ile de Cuba. [Vol. II] Paris. 1845. Plantes cellulaires. Pages xiv + 349. In: J.B. Hombron and H. Jacquinot (eds.), Voyage au Péle Sud et dans l'Océanie sur les corvettes l'Astrolabe et la Zelée, exécute ... pendant les années 1837—1538—1839-1840, sous le commandement de M.J. Dumont- d'Urville. Atlas: Botanique: Cryptogamie. Vol. 1. Paris. 1846. Ordo I. Phyceae Fries. Pages ii + 197. In: M. C. Durieu de Maisonneuve (ed.), Exploration scientifique de l'Algérie pendant les années 1840, 1841, 1842. Sciences naturelles: Botanique, Partie 1. Cryptogamie. Imprimerie Royale: Paris. Murray G. 1887. Catalogue of Ceylon algae in the Herbarium of the British Museum. Annals & Magazine of Natural History, series 5, 20: 21-44. 204 Nageli, C. 1862. Beitrage zur Morphologie und Systematik der Ceramiaceae. Sitzungsberichte der K@oniglichen Bayerischen Akademie der Wissenschaften zu Miinchen 1861(2): 297-415. Norris, J. N., and K. E. Bucher 1982. Marine algae and seagrasses from Carrie Bow Cay, Belize. Smithsonian Contributions to the Marine Sciences, No. 12: 167-223. ‘ Oliveira, E. C. de 1977. Algas marinhas bentonicas do Brasil. Universidade de Sao Paulo, Instituto de Biociéncias: Sao Paulo, Brazil. iv + 407 pp. [Thesis] Olsen, J. L., and J. A. West 1988. Ventricaria (Siphoncladales-Cladophorales complex, Chlorophyta), a new genus for Valonia ventricosa. Phycologia 27: 103-108. Ostenfeld, C. H. 1902. Hydrocharitaceae, Lemnaceae, Pontederiaceae, Potamogetonaceae, Gentianaceae (Limnanthemum), Nymphaeaceae. Botanisk Tidsskrift 24: 260-263. Page, J. Z. 1970. Existence of a Derbesia phase in the life history of Halicystis osterhoutii Blinks and Blinks. Journal of Phycology 6: 375-380. Papentuss, G. F. 1968. Notes on South African marine algae. V. Journal of South African Botany 34: 267-287. Parke, M. W., and P. S. Dixon 1976. Check-list of British marine algae—third revision. Journal of the Marine Biological Association of the United Kingdom 56: 527-594. Rabenhorst, L. 1847. Deutschlands Kryptogamen-Flora ... Zweiter Band. Zweite Abteilung: Algen. Leipzig. xIx + 216 pp. Reinke, J. 1889. Algenflorda der westlichen Ostsee deutschen Anthiels. Eine systemaisch- pflanzen- geographische Studie. Bericht der Kommission zur Wissenschaftlichen Untersuchungen der Deutschen Meere in Kiel 6: iii—xi, 1-101. Rodriguez de Rios, N., and Y. Saito 1982. Observaciones sobre el género Laurencia en Venezuela. |. Laurencia intermedia Yamada y Laurencia corallopsis (Montagne) Howe. Ernstia 11: 1-16. Roth, A. W. 1806. Catalecta botanica ... Fasc. 3. Lipsiae [Leipzig]. [vii] + 350 + [9] pp. Riitzler, K., and I. G. Macintyre (eds.) 1982. The habitat distribution and community structure of the barrier reef complex at Carrie Bow Cay, Belize. Pages 9-45. Jn K. Ruetzler and I. G. Macintyre [eds.], Zhe Atlantic barrier reef ecosystem at Carrie Bow Cay, Belize, I. Structure and communities. Smithsonian Contributions to Marine Science, No. 12. Schmitz, F. 1893. Die Gattung Lophothalia J. Ag. Berichte der Deutschen Botanischen Gesellschaft \1: 212-232. Schnetter, R., 1. Hérnig, and G. Weber-Peukert 1987. Taxonomy of some North Atlantic Dictyota species (Phaeophyta). Hydrobiologia 151/152: 193-197. Searles, R. B., and C. W. Schneider 1989. New genera and species of Ceramiaceae (Rhodophyta) from the southeastern United States. Journal of Phycology 25: 731-740. 205 Silva, P. C. 1960. Codium (Chlorophyta) in the tropical western Atlantic. Nova Hedwigia |: 497-536. Silva, P. C., P. W. Basson, and R. L. Moe 1996. Catalogue of the Benthic Marine Algae of the Indian Ocean. University of California Publications in Botany 79: 1-1259. Silva, P. C., E. G. Mefiez, and R. L. Moe 1987. Catalog of the benthic marine algae of the Philippines. Smithsonian Contributions to the Marine Sciences 27. iv + 179 pp. Smith, J. E. 1790-1814. English botany ... 36 volumes. London. 2592 plates. Stephenson, T. A. 1944. The constitution of the intertidal fauna and flora of South Africa. Part Il. Annals of the Natal Museum 10: 261-358, pls. XII-XIV. Taylor, P. R., M. M. Littler, and D. S. Littler 1986. Escapes from herbivory in relation to the structure of mangrove island macroalgal communities. Oecologia 69: 481-490. Taylor, W. R. 1928. The marine algae of Florida with special reference to the Dry Tortugas. Publications of the Carnegie Institution of Washington 379. [v] + 220 pp. 1960. Marine algae of the eastern tropical and subtropical coasts of the Americas. University of Michigan Press: Ann Arbor. xi + [ili] + 870 pp. Taylor, W. R., and I. A. Abbott 1973. Anew species of Botryocladia from the West Indies. British Phycological Journal 8: 409-412. Tilden, J. E. 1894-1909. American algae. Minneapolis. Centuries 1-7, nos. 1-650. [Exsiccata with printed labels. ] 1910. Minnesota algae, Vol I. Minneapolis. Iv + 319 pp., 20 pls. Tsuda, R. T., and C. J. Dawes 1974. Preliminary checklist of the marine benthic plants from Glover's Reef, British Honduras. Atoll Research Bulletin 173. 13 pp. Turner, D. 1811-1819. Fuci... Vol. 4. London. [iti] + 153 + [2] + 7 pp., 197-258 pls. Vahl, M. 1802. Endeel kryptogamiske Planter fra St.-Croix. Skrifter af Naturhistorie-Selskabet, Kiobenhavn [Copenhagen]. 5(2): 29-47. Williams, L. G. 1951. Algae of the black rocks. Journal of the Elisha Mitchell Scientific Society 67: 149-159. Withering, W. 1776. A botanical arrangement of all the vegetables naturally growing in Great Britian ... M. Swinney: Birmingham. [1]—xvill, xvii-xcvi + 838 pp., XII pls. Womersley, H. B. S. 1978. Southern Australian species of Ceramium Roth (Rhodophyta). Australian Journal of Marine & Freshwater Research 29: 205-257. Woodward, T. J. 1794. Description of Fucus dasyphyllus. Transactions of the Linnean Society of London 2: 239-241. 1797. Observations upon the generic character of Ulva, with descriptions of some new species. Transactions of the Linnean Society of London 3: 46-58. 206 Wulfen, F. X. 1803. Cryptogama aquatica. Archiv fiir die Botanik 3: |-64. Wynne, M. J. 1998. A checklist of benthic marine algae of the tropical and subtropical western Atlantic: First revision. Nova Hedwigia 116: iii + 1-155. ATOLL RESEARCH BULLETIN NO. 475 EPIPHYTIC FORAMINIFERA OF THE PELICAN CAYS, BELIZE:DIVERSITY AND DISTRIBUTION BY SUSAN L. RICHARDSON ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 Mexico Resriiens Sea ‘Northeast Cay Bird Cays* i Guatemala Hondurac a e ve : =n ie Sal ‘Ridge Cay Co Cat Cay: : : — PC-96-G iPC-96-C3 eae ee PC-96-C1 N : ake rer a te 7 Pond C Mee i PC-96-A Little Cat Ca : PX Pond G9 Ags : : PC-96-F —} : so 7 Island : aa numerous coral heads SS 0.5 km te Figure 1. Map showing locations of sampling sites (August 1996) in the Pelican Cays, Belize. EPIPHYTIC FORAMINIFERA OF THE PELICAN CAYS, BELIZE; DIVERSITY AND DISTRIBUTION BY. SUSAN L. RICHARDSON' ABSTRACT The diversity and distribution of epiphytic foraminifera living on the seagrass Thalassia testudinum were surveyed at six localities in the Pelican Cays, Belize. A total of seven species, two of them new, were identified from these sites. Estimates of standing stock range from 6.35 x 103 to 6.90 x 10* individuals/m? of the seafloor, and population densities range from 13.60 to 80.81 individuals/100 cm? of leaf surface area. The faunal assemblages are characterized by low species richness (S = 3 to 6), high dominance (37.91 to 89.91%), and moderate evenness (E = 0.42 to 0.80). A SHE analysis (Buzas and Hayek, 1996) performed for the Pelican Cays data indicates that the distribution of epiphytic foraminifera on Thalassia most closely resembles a log-series pattern (Fisher et al., 1943). INTRODUCTION As organisms, benthic foraminifera form an integral component of seagrass communities in the tropical Western Atlantic region, living both in the sediments (Bock, 1967, 1971; Buzas et al., 1977) and as epiphytes on blades of seagrass (Brasier, 1975 a, 1975b; Steinker and Steinker, 1976; Steinker and Rayner, 1981; Martin, 1986; Waszczak and Steinker, 1987; Martin and Wright, 1988). Previous studies of Belizean foraminiferal faunas have focused exclusively on the sediment-dwelling assemblages (Cebulski, 1969; Wantland, 1975). Wantland (1975, p. 358) observed the highest diversities and abundances in monospecific stands of the seagrass Thalassia testudinum and therefore speculated that most benthic foraminiferal inhabitants of "shallow back- reef environments live attached to plants and other floral and faunal elements above the sediment surface." The objective of this study was to survey the diversity and distribution of the foraminiferal species living on Thalassia testudinum in the Pelican Cays, Belize, Central America (Fig. 1). This paper presents the results of preliminary field collections and observations that took place in August 1996. ‘Department of Ecology and Evolutionary Biology, Yale University, New Haven, CT, USA. Present Address: Department of Geology and Geophysics, MS-08, Woods Hole Oceanographic Institution, Woods Hole, MA 02540, USA. 210 MATERIAL AND METHODS Samples of the seagrass Thalassia testudinum Banks ex KG6nig (Fig.2) were collected by snorkeling from seven localities in the Pelican Cays: off Cat Cay, Pond A of Cat Cay, three locations in Pond C of Manatee Cay, Pond J of Little Cat Cay, and Pond E of Fisherman's Cay (Fig. 1, Appendix 1). Samples were collected by removing all shoot bundles and attached seagrass blades from a 10-by-20 cm quadrat. They were then transported to the lab in a cooler, fixed in 4-5% formaldehyde in seawater, and transferred to 70% EtOH for storage. Live individuals still attached to their seagrass substrate were examined in the wet lab on Carrie Bow Cay using a binocular microscope (Wild M3). At Pond A, J, and C sites, shoot densities of Thalassia were estimated by counting all shoots in two 25-by-25 cm quadrats, and the mean value was used to calculate the shoot densities per m? of seafloor. At the Pond E site, shoot densities were estimated by counting all shoots in a single 10-by-20 cm quadrat. Leaf area indices (LAIs) were calculated for each site by measuring all Thalassia blades collected from two 10-by-20 cm quadrats and using the mean value as an estimate of the leaf surface area available for settlement by epiphytic organisms per m? of seafloor. Epiphyte load was determined from the average dry weights of leaves and epiphytes removed from all Thalassia blades collected from two 10-by-20 cm quadrats. Epiphytes were scraped from both sides of each leaf with a razor blade, and leaves and epiphyte scrapings were dried for 8 hrs. at 105°C prior to weighing. For the purposes of this study, an epiphyte is defined as "any organism that lives upon a plant and completes its production while it is still attached to that plant. This definition includes the coralline red algae but excludes mobile gastropods and benthic foraminifera which are able to move between leaves and thus are likely to produce for larger periods of time " (Frankovich and Zieman, 1994, p. 682). This definition corresponds to Langer's (1993) category of "permanently attached" epiphytic foraminifera, but excludes species he categorized as "temporarily attached" and "motile." All epiphytic foraminifera were within a single 10-by-20 cm quadrat of Thalassia blades for each locality. Examination of fresh material in the laboratory on Carrie Bow Cay showed that all specimens still attached to the leaf blades contained cytoplasm and were alive. Live specimens were recognized by evidence of pseudopodial activity, feeding cysts and cytoplasmic coloration. The original cytoplasmic coloration was also preserved in samples that had been initially fixed in 5% formaldehyde, then transferred to 70% EtOH. RESULTS The estimated number of Thalassia shoots per m? of seafloor varied from site to site in the Pelican Cays (Table 1). The highest shoot densities were seen in Pond A (928 + 32 shoots/m? seafloor), while the lowest densities were recorded just outside this pond, off the western side of Cat Cay (424 + 56 shoots/m? seafloor). Ponds C and E yielded similar estimates of shoot density, 616 + 56 and 600 shoots/m’ seafloor, respectively. Leaf area indices calculated for each site appear to correlate with shoot densities in general, ranging from a low value of 1.99 at Cat Cay to a high value of 4.01 at Pond J. Observed values of epiphyte and seagrass dry weights appear to be positively correlated with shoot densities for each site. Seagrass dry weights range from 37.18 + 17.13 gdw/m’ seafloor at the Cat Cay site to 77.36 + 17.18 gdw/m’ seafloor at the Pond J site, ZA Figure 2. Examples of epiphytic foraminifera. A) Schematic illustration of seagrass Thalassia testudinum with detail of a blade showing attached epiphytic foraminifera (M.E. Parish after I.C. Feller). B) Scanning electron photomicrograph of Belizeanella candeiana (scale = 100 jm). 212 while epiphyte dry weights range from 3.48 + 0.48 gdw/m* seafloor to 7.20 + 2.05 gdw/m? seafloor at these same sites, respectively. Values of epiphyte load were calculated from the measured dry weights of the epiphytes and seagrass at each site in accordance with the method of Tomasko and Lapointe (1991). The lowest value of epiphyte load was observed at Pond A (7.56 + 6.06%), the highest value at Pond A (10.85 + 5.61%). Intermediate values of epiphyte load were calculated for the sites at the entrance to Pond C (8.08 + 3.40%) and Pond J (8.41 + 0.52%). Table 1. Seagrass and epiphyte data from Pelican Cays sample locations. Pond C (entrance) Parameter* Pond J Mean shoot density (# shoots/ m? seafloor)? Cat Cay Pond A 424+56 928+ 32 616 + 56 808 + 152 600 Mean leaf area index 1.99 + S) AST 3e 3.1340.26 4.01+0.90 2.59 (LAI) (m? leaf 0.78 (0.70 surface area/m? seafloor) Mean epiphyte 3.48 + 5.00" 3-7 5.15 120 205 — biomass (gdw/m? 0.48 2.58 seafloor) Mean seagrass biomass (gdw/m? seafloor) Epiphyte load (%)* 37.18 + TOM5 3 56.40 + T7362 WEN = 703 a5 3.4] T50e 8.08+3.40 8.41+0.52 = 6.06 “Mean values calculated from two quadrats + deviation from mean. Shoot density calculated from a single 20-by-20-cm quadrat; LAI value calculated from a single 10-by-20-cm quadrat collected for epiphytic foraminiferal census. “Weight epiphytes(weight epiphytes + weight seagrass) x 100. Standing stock estimates of epiphytic foraminiferal populations were calculated from the total number of individuals counted in each quadrat (Tables 2 and 3). The highest standing stock was observed on the ridge at Pond C (6.90 X 10? individuals/m’seafloor); however, just inside the lagoon, the standing stock drops by an order of magnitude (7.85 X 10° individuals/m’seafloor). Relatively high estimates, 1.29 X 10* and 2.09 X 10% individuals/m? seafloor, respectively, were obtained from both the Pond A and Pond E sites. The lowest standing stock estimates were calculated for the Cat Cay (7.65 X 10° individuals/m? seafloor) and Pond J sites (6.35 X 10° individuals/m? seafloor). Table 2. Number of individuals of epiphytic Foraminifera per sample from Pelican Cays locations. Pond C Pond C Pond A (ridge) (lagoon) PondJ Pond E Species Cat Cay Belizeanella candeiana ] 0 0 0 0 0 Cornuspira planorbis 0 0 2 0 0 0 Cornuspiramia antillarum 58 194 549 0 Sil 48 Hemidiscella palabunda 0 2 10 0 0 Tridia n. sp. 36 2 47] 141 14 62 Planorbulina acervalis 56 339 6 82 307 Rhizonubecula n. sp. Total Table 3. Standing stock of Foraminifera from Pelican Cays sample locations (# individuals/m? seafloor). PondC Pond C Species Cat Cay PondA (ridge) (lagoon) PondJ PondE Belizeanella candeiana 5.00 x 10! 0 0 0 0 0 Cornuspira planorbis 0 0 1.00 x 10? 0 0 0 Cornuspiramia antillarum | 2.90 x 10? 9.70 x 10° 2.75 x 104 0 1.55x 10° 2.40 x 10° Hemidiscella palabunda 0 5.00x 10! 1.00x 10? 5.00 x 10° 0 0 1.80x 10° 1.05x 10? 2.36x10* 7.05x 10°? 7.00x 10? 3.10x 10° 2.80x 10? 2.05x 10? 1.70x10* 3.00x 10? 4.10x 10% 1.53 x 104 1.00x 10° 5.00x 10! 8.50 x 10? 0 0 1.00 x 10? 7.65x 10? 1.29x10* 6.90x10* 7.85x 10? 6.35x 10’ 2.09 x 104 Tridia n. sp. Planorbulina acervalis Rhizonubecula n. sp. Total Population densities of epiphytic foraminifera on Thalassia blades were calculated from the measured leaf surface area per quadrat and the total number of individuals counted in each sample (Table 4). The highest densities were observed in the samples collected from Pond E (80.81 individuals/100 cm? of leaf surface area) and the ridge in Pond C, Manatee Cay (69.38 individuals/100 cm? of leaf surface area). Intermediate densities were seen at Cat Cay (40.87 individuals/100 cm? of leaf surface area) and Pond A (36.58 individuals/100 cm? of leaf surface area), while the lowest densities were observed in samples collected from Pond J (16.97 individuals/100 cm? of leaf surface area) and Pond C (13.60 individuals/100 cm? of leaf surface area). 214 Table 4. Relative density of Foraminifera at Pelican Cays sample locations (# individuals/100m? leaf surface area). PondC Pond C (ridge) (lagoon) PondJ Pond E Species Cat Cay Pond A Belizeanella candeiana Cornuspira planorbis 0 0 0.10 0 0 0 Cornuspiramia antillarum | 15.49 27.51 27.60 0 4.14 9.26 Hemidiscella palabunda 0 0.14 0.10 0.86 0 0 Tridia n.sp. 9.62 2.98 23.68 12.22 1.87 11.96 Planorbulina acervalis 14.96 5.8] 17.04 0.52 10.96 SII Rhizonubecula n.sp. Total Thalassia provides substrate for a variety of organisms, such as encrusting coralline algae, filamentous algae, spirorbid polychaetes, bryozoans, hydroids, anemones, sponges, and molluscan egg cases (Table 5). Encrusting coralline algae are among the most conspicuous members of the epiphytic community and heavily encrust the leaf margins and distal portion of Thalassia blades at both Pond A and Pond C ridge sites. The diversity of nonforaminiferal epiphytes, including extremely abundant spirorbids, was highest (N = 340) in the sample collected from the ridge in Pond C. Table 5. Distribution of nonforaminiferal epiphytes at Pelican Cays sample locations. Pond C Pond C Species Cat Cay PondA (ridge) (lagoon) PondJ PondE Light- | Moderate- Moderate- Light- Light- Coralline algae" moderate heavy heavy Absent moderate moderate Filamentous algae” 0 0 a 0 a 0 Sponges 0 0 I 2 0 0 Hydroids a 0 0 0 0 0 Anemones | 0 0 0 0 0 Bryozoans 0 l 2 0 0 0 Spirorbid worms 0 | 340 12 15 1] Egg masses 0 0 2 0 0 0 “Coralline algae not identified. However, Littler and Littler (1997) report the following taxa occurring as epiphytes on Thalassia in the Pelican Cays: Fosliella farinosa (Lamouroux), Pheumophyllum fragile Kiitzing, and Titanoderma pustulatum (Lamouroux). >Filamentous algae not identified. However, Littler and Littler (1997) record the following taxa occurring as epiphytes on Thalassia in the Pelican Cays: Champia parvula (C. Agardh) var. parvula, Ceramium flaccidum (Kiitzing), Wrangelia penicillata (C. Agardh), Polysiphonia flaccidissima Hollenberg, P. scopulorum Harvey, Feldmannia indica (Sonder in Zollinger), Rosenvigea sanctae- crucis Borgesen, and Sphacelaria tribuloides Meneghini. +Phaeophyte tentatively identified as Dictyota sp. attached to Thalassia blades at this locality. 215 A low-diversity assemblage composed of the following seven species comprised the total community of epiphytic foraminifera living on the seagrass Thalassia testudinum in the Pelican Cays: Belizeanella candeiana (d’ Orbigny), Cornuspira planorbis Schultze, Cornuspiramia antillarum (Cushman), Hemidiscella palabunda (Bock), Iridia n. sp., Planorbulina acervalis Brady, and Rhizonubecula n. sp. The species Cornuspiramia antillarum has previously been reported from Belize, off Carrie Bow Cay (Richardson, 1996), and has been cited as a minor component of the shallow-water foraminiferal faunas in other regions of the tropical Western Atlantic (Cushman, 1922, 1929; Bérmudez, 1935; Hofker, 1964, 1971, 1976; Brasier 1975a; Manning, 1985). The two species recognized as new, /ridia n. sp. and Rhizonubecula n. sp., also occur as seagrass epiphytes off Carrie Bow Cay and the Twin Cays (Richardson, 1996). The remaining four species—B. candeiana, C. planorbis, H. palabunda, and P. acervalis—have been recorded from sediments of the Belizean shelf by Cebulski (1969) and Wantland (1975) (see Appendix 2 for synonomies). Epiphytic foraminifera are known to attach to a variety of substrates other than seagrasses (Brasier, 1975a, 1975b; Langer, 1988, 1993). Samples of Turbinaria sp. from Pond A contained a few juvenile specimens of P. acervalis attached to the blades and a relatively dense growth of epiphytes on the stalk (including filamentous algae, erect bryozoans, and molluscan egg masses), but few foraminifera (2-3 Jridia n. sp.). Examination of Halimeda sp. from Pond C revealed only minor epiphytes: filamentous algae and a few specimens of /ridia n. sp. At Pond J, examination of several individuals of Penicillus sp. yielded a few spirorbids and two foraminifera (1 P. acervalis and | C. antillarum); however, dense epiphytic growth was observed to cover Halimeda sp. collected from the same locality. A single individual of Halimeda sp. contained numerous specimens of adult P. acervalis, several C. antillarum, and a few Iridia n. sp., in addition to moderate to heavy encrustation by calcareous algae, spirorbid worms, Dictyota sp., and a few sponges and anemones. The relative abundance of foraminiferal species living on Thalassia varies, with different species dominating in different proportions at each site (Tables 6 and 7). Species dominance is high, ranging from 37.91% in the Pond A sample to 89.81% in the Pond C sample. Cornuspiramia antillarum is the most abundant species at Cat Cay, Pond A, and Pond C ridge sites, and the second most abundant species at Pond J. C. antillarum was not found in the sample collected in Pond C because encrusting coralline algae are not present on seagrasses at this site. Cornuspiramia antillarum has been observed to preferentially encrust calcareous substrate such as coralline algae or shell fragments (S. Richardson, unpublished observations). Planorbulina acervalis 1s the most abundant species at the Pond J and Pond E sites, the second most abundant species at the Cat Cay and Pond A sites, and the third most abundant species in the Pond C ridge sample. /ridia sp. accounts for the most abundant species at the Pond C lagoon site, the second most abundant species at the Pond C ridge and Pond E sites, and the third most abundant species in Pond J, off Cat Cay, in Pond A. Hemidiscella palabunda was the second most abundant species in the sample from Pond C, but comprised less than 1% of assemblages from the Pond C ridge and Pond A samples. Rhizonubecula n. sp. was recorded in abundances of less than 2% at all sites, except Ponds J and C, where it was not found at all. Belizeanella candeiana and Cornuspira planorbis were recorded in abundances of less than 1% at only a single site each, Cat Cay and the ridge in Pond C, respectively. 216 Table 6. Relative abundance of foraminiferan species at Pelican Cays sample locations (percent). ; PondC Pond C (ridge) (lagoon) PondJ Pond E Species Cat Cay Pond A Belizeanella candeiana 0.65 0 0 0 0 0 Cornuspira planorbis 0 0 0.15 0 0 0 Cornuspiramia antillarum| 37.91 7519. 39.78 0 24.41 11.46 Hemidiscella palabunda 0 0.39 0.15 6.37 0 0 Tridia n. sp. 23253 8.14 34.13 89.81 11.02 14.80 Planorbulina acervalis 36.60 15.89 24.56 3.82 64.57 7326 Rhizonubecula n. sp. esl 0.39 123 0 0 0.48 Total 100 100 100 100 100 100 Table 7. Rank abundance of foraminiferan species at Pelican Cays sample locations making up more than 5% of assemblage. Pond C Pond C Cat Cay Pond A (ridge) (lagoon) Pond J Pond E Cornuspiramia Cornuspiramia Cornuspiramia Iridian. sp. Planorbulina — Planorbulina antillarum antillarum antillarum acervalis acervalis Planorbulina — Planorbulina Iridian. sp. Hemidiscella Cornuspiramia Iridia n. sp. acervalis acervalis palabunda — antillarum Tridia n. sp. Iridia n. sp. Planorbulina Iridia n. sp. Cornuspiramia antillarum Rank NO Lvs) acervalis The highest species richness (S) was observed in the sample collected from the ridge in Pond C (S = 6), followed by the sites off Cat Cay (S=5) and in Pond A (S = 5). The least speciose sites were observed to be Pond E (S=4), Pond J (S = 3), and inside Pond C (S = 3). In addition to species richness (S), values of the Shannon information function (H), and the Buzas and Gibson (1969) evenness function (E) were calculated for each sample individually (Table 8). These indices together can be used as measures of species diversity (S) and species equitability or dominance (E) (Hayek and Buzas, 1997). In addition to H, the equitability measure J was also calculated for each sample, because this measure is considered to be less dependent on S when the species number is less than 10 (Sheldon, 1969; Gibson and Buzas, 1973). In the Pelican Cays samples, values of H range from 0.40 in Pond C to 1.17 off Cat Cay, and values of E range from 0.42 in Pond A, to 0.80 off Pond J. The values of J exhibit a range similar to E's, from 0.36 in Pond C to 0.79 off Cat Cay (Table 8). A SHE analysis (Buzas and Hayek, 1996; Hayek and Buzas, 1997) was performed for the Pelican Cays data. This procedure consists of calculating the values of H and E for cumulative quadrats, and then determining how these values change as a function of the number of individuals (N) (Fig. 2, Table 9). Results from the SHE analysis indicate that the distribution of epiphytic foraminifera on Thalassia in the Pelican Cays most closely resembles a log-series pattern (Fisher et al., 1943). As discussed by Buzas and Hayek (1996) and Hayek and Buzas 217 (1997), this pattern is one in which values of H remain relatively constant with increasing N. Table 8. Summary of data on distribution of Foraminifera at Pelican Cays sample locations. PondC PondC Parameter Cat Cay PondA_ (ridge) (lagoon) PondJ PondE S (# species/sample) 5 5 6 3 3 4 N (# specimens/sample) 153 258 1380 157 127 419 H (Information function)* iY 0.75 NS 0.40 0.87 0.78 E (Evenness measure)? 0.64 0.42 0.53 0.50 0.80 0.54 J (Equitability measure)° 0.72 0.47 0.64 0.36 0.79 0.56 # leaves surveyed 34 50 47 25 44 33 Length: width (average for all] 6.82 ) 15.71 19.19 eX) 1BFSil leaves) Total leaf surface area (cm’) Density (# forams/100 cm? leaf ) Standing stock (# forams/m? seafloor) “Shannon information function: H = - ¥ p, In (p,) (Hayek and Buzas, 1997). >’Buzas and Gibson (1969) measure of equitability or evenness: E = e"/S (Hayek and Buzas, 1997). “Equitability measure: J = H/In S (Pielou, 1966). 374.34 705.30 1988.96 1154.18 748.38 518.50 40.87 36.58 69.38 “13.60 16.97 80.81 7.65 x 10° 1.29x 10* 6.90x10* 7.85x 10? 6.35x 10°? 2.09 x 104 y, 1 Fema eenmiind haa Akiv 48 InE —-6 InE/Ins —e H 4 e 0 -e InS 0 1000 2000 3000 N Figure 2. SHE analysis plot for Pelican Cays epiphytic Foraminifera data (refer to Table 9). 218 Table 9. SHE analysis for Pelican Cays data on distribution of Foraminifera at Pelican Cays sample locations (cf. Fig. 2). Quadrat S 1 127 3 0.8698 0.7954 -0.2289° 1.0986 2 284 4 105311 O71G7, -les3si 1.3863 = -0.2403 3 703 5 1.0095 0.5488 -0.6000 1.6094 = -0.3728 4 856 6 1.0683 0.4851 -0.7234 1.7918 -0.4037 6 2236 i 1.1768 0.4634 -0.7692 1.9459 — -0.3953 Note: N = cumulative number of individuals; S = cumulative number of species; H = - )’ p; In (p;) (Hayek and Buzas, 1997); E = el'/S (Hayek and Buzas, 1997; Buzas and Gibson, 1969). Samples were added to SHE analysis in order of increasing species richness. Quadrat # 5 (Pond A) was dropped from the analysis because it resulted in an anomalously high value of E. DISCUSSION Thalassia testudinum Banks ex K6nig is the dominant seagrass in the Caribbean and grows in extensive meadows in shallow waters down to 20 m (den Hartog, 1970; Littler et al., 1989; Norris and Bucher, 1982; Phillips and Mefiez, 1988). The vegetative morphology of Thalassia consists of horizontal rhizomes (= long shoots) that grow beneath the sediment, and branching from them are lateral, erect leaf-bearing short shoots (Tomlinson and Vargo, 1966; Tomlinson, 1974). Leaves grow from the base, and new leaves are produced from the center of the leaf bundles (Tomlinson and Vargo, 1966; Tomlinson, 1974). The shoot densities of Thalassia in the Pelican Cays are within the range of shoot densities measured by the author for Thalassia in 0.5-m water depths off Carrie Bow Cay (range = 552-1,160 shoots/m’ seafloor) (S. Richardson, unpublished observations), but somewhat higher than previous estimates for several localities in Belize (117-404 shoots/m? seafloor) (Tomasko and Lapointe, 1991). Estimates of seagrass biomass in the Pelican Cays were observed to be slightly lower than off Carrie Bow Cay (range = 73.50-108.85 gdw/m? seafloor) (S. Richardson, unpublished observations), but higher than previously published estimates for other localities in Belize (range = 17.30-49.20 gdw/m’ seafloor) (Tomasko and Lapointe, 1991). Estimates of epiphyte biomass (range = 12.70-41.25 gdw/m? seafloor) and epiphyte load (range = 10.73-31.26%) obtained for Thalassia off Carrie Bow Cay (S. Richardson, unpublished observations) were found to be considerably higher than the values obtained for Thalassia in the Pelican Cays (Table 1). The most significant environmental factor known to influence the biomass of seagrass epiphytes is the nutrient content of the water column (Borowitzka and Lethbridge, 1989). High levels of dissolved nutrients in the water column (e.g., ammonium, nitrite plus nitrate, dissolved inorganic nitrogen, and soluble reactive phosphate) have been shown to be correlated with higher epiphyte loads (Tomasko and Lapointe, 1991; Frankovich and Fourqurean, 1997). For example, Tomasko and Lapointe (1991) report levels of dissolved nutrients that are 6 to 25 times higher in the water column off Big Pine Cay, Florida, than off Carrie Bow Cay, Belize. Likewise, epiphyte loads off Big Pine Cay, Florida, are three times higher than at Carrie Bow Cay, and 4 to 6 times 219 higher than the values calculated for the Pelican Cays (Tomasko and Lapointe, 1991). The oligotrophic waters of the Pelican Cays represent a pristine environment isolated from anthropogenic pollution (Littler and Littler, 1997) and, correspondingly, low values of epiphyte load have been recorded for these sites (Table 1). The maximum standing stocks calculated for epiphytic foraminifera living on Thalassia testudinum in the Pelican Cays (Tables 3, 8) are comparable to the values obtained by the author for Carrie Bow Cay (6.82 x 10* individuals/m’ seafloor) and are similar to published estimates from other regions in the Western Atlantic and world's oceans. According to Erskian’s (1972) estimates, population densities of Planorbulina sp. and Sorites sp. on Thalassia in Discovery Bay, Jamaica, exceed 6.0 x 10° individuals/m? seafloor and 1.2 x 10° individuals/m? seafloor, respectively. In Barbuda, between 1.24 x 10* and 2.07 x 10° epiphytic foraminifera/m’ seafloor live attached to various phytal substrates in depths less than 2 m (Brasier, 1975a). In the Gulf of Elat, Red Sea, an estimated 1.54 x 10° epiphytic foraminifera/m? seafloor have been recorded living on the both the leaves and rhizomes of Halophila stipulacea collected from 20 m (Faber, 1991). In the Mediterranean, off Banyuls-sur-Mer, France, the standing stock of epiphytic foraminifera living on Posidonia oceanica increases with increasing water depth, with an estimated 3.0 x 10* individuals/m? seafloor reported at 5 m and 1.7 x 10° individuals/m? seafloor at 20 m (Vénec-Peyré and Le Calvez, 1988). The densities of epiphytic foraminifera observed living on Thalassia blades in the Pelican Cays (Table 4) are similar to but somewhat higher than the densities reported by Wilson (1998) for epiphytic foraminifera living on Thalassia (19.05 individuals/100 cm? leaf surface area) and Syringodium (17.65 individuals/100 cm? leaf surface area) in Cockleshell Bay, St. Kitts. Brasier (1975a), however, has reported exceedingly high densities (4,000-8,333 individuals/100 cm? leaf surface area) of epiphytic foraminifera living on Thalassia collected off Barbuda. Lewis and Hollingworth (1982) recorded total densities ranging from 1.69 to 1757 individuals/100 cm? leaf surface area for all epiphytic organisms (excluding foraminifera) encrusting Thalassia blades collected from a variety of localities off Barbados. Seagrass epiphytes must settle, grow, and reproduce within the life span of an individual blade, and they exhibit the rapid growth and reproductive rates characteristic of opportunistic species (Keough, 1986; Borowitzka and Lethbridge, 1989; Dirnberger, 1990, 1993, 1994; Kaehler and Hughes, 1992). Few details are known, however, about the life history traits of most species of epiphytic foraminifera. Cushman (1922, p. 59) documented a rapid growth rate for the species Cornuspiramia antillarum, observing that it was "one of the first organisms to be attached to the leaf." Previous authors have assumed an annual life span for Planorbulina acervalis and other epiphytic species (Le Calvez, 1936, 1938; Lutze and Wefer, 1980; Zohary et al., 1980; Hallock et al., 1986; Langer, 1988, 1993; Vénec-Peyré and Le Calvez, 1989; Hottinger, 1990). Recently, however, several specimens of P. acervalis, as well as Jridia n.sp., were found to reproduce by multiple fission and to still contain juveniles within the parental test (S. Richardson, unpublished observations), indicating that the generation time of these species falls within the life span of Thalassia blades. The low total species richness of the epiphytic fauna in the Pelican Cays (S = 7) contrasts sharply with the higher diversities that characterize epiphytic foraminiferal faunas described from other localities of the tropical Western Atlantic region. Brasier (1975a) identified a total of 49 species from various phytal substrates off Barbuda. Martin and Wright (1988) recorded 69 foraminiferal species living on Thalassia in the back-reef lagoon off Key Largo, Florida. Bock 220 (1969) reported 66 species epiphytic on Thalassia off Big Pine Key, Florida; 18 occurring in abundances >1% and 10 abundant throughout the year. Waszczak and Steinker (1987) recorded a total of 106 species of epiphytic foraminifera living on a variety of algal and seagrass substrates off Big Pine Key, Florida. These higher species diversities reflect, in part, the inclusion of mobile epiphytic species in the tallies of previous studies. Recently, Wilson (1998) described an assemblage of only 11 species of epiphytic foraminifera living on Thalassia testudinum and Syringodium filiforme in Cockleshell Bay, St. Kitts. High-diversity epiphyte communities have been correlated with longer-lived phytal substrates (Borowitzka and Lethbridge, 1989; Langer, 1988, 1993; Hottinger, 1990). In Belize, the life span of individual leaves is short (35.3 to 42.7 days) and blade turnover rates are relatively high (2.34 to 2.83% a day) (Tomasko and Lapointe, 1991; Koltes et al., in press). As growth rates of Thalassia testudinum are relatively uniform throughout the Caribbean (Patriquin, 1973; Zieman, 1974; Zieman and Wetzel, 1980), the relatively low species diversity of the Pelican Cays fauna must be related to other factors. Shallow-water tropical marine environments are generally characterized by high species richness, a trend that has been documented in benthic marine invertebrates, as well as benthic foraminifera (Fisher, 1960; Sanders, 1968, 1969; Buzas, 1972). Low species diversities are believed to characterize "physically controlled communities" in which the constituent organisms are subject to fluctuating environmental conditions and high physiological stress (Sanders, 1968, 1969). For example, Gibson and Buzas (1973) found lower species richness to characterize benthic foraminiferal faunas in areas subject to greater physical stress. And Gibson and Hill (1992) found low species richness (2-11 species), coupled with high dominance (30-95%), and low to moderate values of evenness (0.20-0.40), to characterize benthic foraminiferal faunas living in highly variable ecological habitats off the east coast of North America. The Pelican Cays fauna exhibit low species richness (3-6 species), high dominance (37.91-89.81%), and moderate species equitability (0.42-0.80) (Table 8); however, the overall environmental regime in this area is relatively constant (Koltes et al., in press). Wilson (1998) also records values of low species richness (5-11), high dominance (66.50-81.50%), and low to moderate equitability (0.30-0.40) for epiphytic foraminifera on Thalassia and Syringodium from Cockleshell Bay, St. Kitts, which suggests that this pattern may be representative of permanent, nonmotile epiphytic communities. One environmental factor that has been shown to affect the species richness and abundance of epiphytic foraminifera is water turbulence (Bock, 1969; Ribes and Gracia, 1991). Martin and Wright (1988) observed a decrease in species richness and increase in species abundances with increasing distance from shore, which they attributed to increased wave, current, and storm activity. Waszczak and Steinker (1987) reported a general increase in species richness with increasing distance from shore but speculated that this trend resulted from the greater stability of the outer-reef environment. The coral-ridge system at the entrance to, and within, Pond C has been observed to affect circulation patterns within the lagoon (Macintyre et al., this volume; Littler and Littler, 1997). The contrast in species composition, relative abundance, and standing stock between epiphytic species living on seagrasses collected from the ridge and from within Pond C indicates that hydrodynamic conditions also play a role in the species composition and distribution of epiphytic faunas of this area. The results from the SHE analysis indicate that the distributional pattern of foraminifera epiphytic on Thalassia in the Pelican Cays is most similar to a log-series pattern (Buzas and 221 Hayek, 1996; Hayek and Buzas, 1997). A log-series pattern is characteristic of communities with relatively few species that are subject to a single, dominant environmental factor (May, 1975). As the log-series model predicts that the greatest number of species will have minimal abundance, an increased sampling effort would be expected to yield a larger number of rare species (Fisher et al., 1943; May, 1975). SUMMARY Estimates of epiphyte load (7.56-10.85%) obtained for Thalassia in the Pelican Cays are much lower than estimates calculated for the area off Carrie Bow Cay and previously published estimates obtained for Thalassia at other sites in Belize. These low values of epiphyte load can be considered an environmental indicator of the pristine water quality in the Pelican Cays, a region removed from the influence of anthropogenic pollution. Increased anthropogenic input to coastal regions has been implicated as the primary factor responsible for the recent worldwide decline in seagrasses. The total epiphytic community of foraminifera living on the seagrass Thalassia testudinum in the Pelican Cays, Belize, is made up of seven species: Belizeanella candeiana (d’Orbigny), Cornuspira planorbis Schultze, Cornuspiramia antillarum (Cushman), Hemidiscella palabunda (Bock), Iridia n. sp., Planorbulina acervalis Brady, and Rhizonubecula n. sp. The two species recognized as new, Jridia n. sp. and Rhizonubecula n. sp., have also been observed as seagrass epiphytes off Carrie Bow Cay and the Twin Cays. The overall pattern of species abundances and distribution in the quadrats sampled is one of low species richness (S = 3-6), high dominance (37.91-89.91%), and moderate evenness (E = 0.42-0.80). This pattern has been previously recognized as characteristic of foraminiferal communities living under stressful conditions (fluctuating salinities and temperatures) in temperate regions, but not of shallow-water tropical reef environments. Results of a SHE analysis indicate that the distribution of foraminiferal species on Thalassia in the Pelican Cays most closely resembles a log-series pattern. ACKNOWLEDGMENTS Fieldwork for this project was supported by a grant from the National Museum of Natural History's Caribbean Coral Reef Ecosystems Program (CCRE Contribution No. 585). I am grateful to Marty Buzas and Jon Moore for their valuable comments on this manuscript, and | would like to extend special thanks to Mike Carpenter and Robyn Spittle for their advice and assistance in the field. REFERENCES Bermudez, P. J. 1935. Foraminiferos de la costa norte de Cuba. Memorias de la Sociedad Cubana de Historia Natural 19(3):129-224, pls. 10-17. Bock, W. D. 1967. Monthly variation in the foraminiferal biofacies on Thalassia and sediment in the Big 222 Pine Key area. Unpublished Ph.D. thesis, University of Miami, Coral Gables, Florida, 243 p. 1968. Two new species of foraminifera from the Florida Keys. Contributions from the Cushman Foundation for Foraminiferal Research 29 (pt. 1):27-29, pl. 4. 1969. Thalassia testudinum, a habitat and means of dispersal for shallow water benthonic foraminifera. Transactions of the Gulf Coast Association of Geological Societies 19:337-340. 1971. A handbook of the benthonic foraminifera of Florida Bay and adjacent waters. In A Symposium of Recent South Florida Foraminifera, edited by J. I. Jones and W. D. Bock, 1-71, pls. 1-24. Miami Geological Society, Memoir 1. Borowitzka, M. A., and R. C. Lethbridge 1989. Seagrass epiphytes. In Biology of Seagrasses: A treatise on the biology of seagrasses with special reference to the Australian region, edited by A. W. D. Larkum, A. J. McComb, and S. A. Shepherd, 2:458-499. Amsterdam and New York: Elsevier, Aquatic Plant Studies. Brady, H. B. 1884. Report on the Foraminifera dredged by H.M.S. Challenger during the years 1873-1876: Reports of the Scientific Results of the Voyage of the H.M.S. Challenger during the years 1873-1876. Zoology 9 (pt. 1) (text):814, pt. 2 (plates), 115 p. Brasier, M. D. 1975a. Ecology of Recent sediment-dwelling and phytal foraminifera from the lagoons of Barbuda, West Indies. Journal of Foraminiferal Research 5(1):42-62. 1975b. The ecology and distribution of Recent foraminifera from the reefs and shoals around Barbuda, West Indies. Journal of Foraminiferal Research 5(3):193-210. Buzas, M. A. 1972. Patterns of species diversity and their explanation. Taxon 21(2/3):275-286. Buzas, M. A., and T. G. Gibson 1969. Species diversity: benthonic foraminifera in western North Atlantic. Science 163:72-75. Buzas, M. A., and L.-A. C. Hayek 1996. Biodiversity resolution: an integrated approach. Biodiversity Research 3:40-43. Buzas, M. A., R. K. Smith, R. K., and K. A. Beem 1977. Ecology and Systematics of Foraminifera in Two Thalassia Habitats, Jamaica, West Indies. Smithsonian Contributions to Paleobiology, v. 31, 139 p. Cebulski, D. E. 1969. Foraminiferal populations and faunas in barrier reef tract and lagoon, British Honduras. American Association of Petroleum Geologists Memoir 11:311-328. Cushman, J. A. 1922. Shallow-water Foraminifera of the Tortugas Region. Department of Marine Biology, Carnegie Institute of Washington, v. 17, no. 311, 85 p., 14 pls. 1929. The Foraminifera of the Atlantic Ocean. Part 6. Miliolidae, Ophthalmidiidae and Fischerinidae. U.S. National Museum Bulletin 104 (pt. 6), 129 p. Dirnberger, J. M. 1990. Benthic determinants of settlement for planktonic larvae: availability of settlement sites for the tube-building polychaete Spirorbis spirillum (Linnaeus) settling onto 223 seagrass blades. Journal of Experimental Marine Biology and Ecology 140:89-105. 1993. Dispersal of larvae with a short planktonic phase in the polychaete Spirorbis spirillum (Linnaeus). Bulletin of Marine Science 52(3):898-910. 1994. Influences of larval settlement location and rate on later growth in a sessile marine invertebrate population (Spirorbis spirillum). Northeast Gulf Science 13(2):65-78. Ehrenberg, C. G. 1839. Ueber die Bildung der Kreidefelsen und des Kreidemergels durch unsichtbar Organismen. Abhandlungen der KGniglichen Akademie der Wissenschaften zu Berlin, 59-147, pls. 1-4. Erskian, M. 1972. Patterns of distribution of foraminifera on Thalassia testudinum. In Marine Studies on the North Coast of Jamaica, edited by G. J. Backus. Atoll Research Bulletin, 152:3. Faber, W. W., Jr. 1991. Distribution and substrate preference of Peneroplis planatus and P. arietinus from the Halophila meadow near Wadi Taba, Eilat, Israel. Journal of Foraminiferal Research 21(3):218-221. Fisher, A. G. 1960. Latitudinal variations in organic diversity. Evolution 14:64-81. Fisher, R. A., A. S. Corbet, and C. B. Williams 1943. The relation between the number of species and the number of individuals in a random sample of an animal population. Journal of Animal Ecology 12:42-58. Frankovich, T. A., and J. W. Fourqurean 1997. Seagrass epiphyte loads along a nutrient availability gradient, Florida Bay, USA. Marine Ecology Progress Series 159:37-50. Frankovich, T. A., and J. C. Zieman 1994. Total epiphyte and epiphytic carbonate production on Thalassia testudinum across Florida Bay. Bulletin of Marine Science 54(3):679-695. Gibson, T. G., and M. A. Buzas 1973. Species diversity: Patterns in modern and Miocene foraminifera of the eastern margin of North America. Geological Society of America Bulletin 84:217-238. Gibson, T. G., and E. E. Hill 1992. Species dominance and equitability: Patterns in Cenozoic foraminifera of eastern North America. Journal of Foraminiferal Research 22(1):34-51. Hallock, P., T. L. Cottey, L. B. Forward, and J. Halas 1986. Population biology and sediment production of Archaias angulatus (Foraminiferida) in Largo Sound, Florida. Journal of Foraminiferal Research 16(1):1-8. Hartog, C. den 1970. The Sea-Grasses of the World. Verhandelingen der Koninklijke Nederlandse Akademie van Wetenschappen, afd. Natuurkunde, v. 59, no. 1. Hayek, L.-A. C., and M. A. Buzas 1997. Surveying Natural Populations. New York: Columbia University Press, 563 p. Hofker, J. 1964. Foraminifera from the tidal zone in the Netherlands Antilles and other West Indian Islands. Studies on the Fauna of Curagao and Other Caribbean Islands 21(83):1-119. 1971. The foraminifera of Piscadera Bay, Curacao. Studies on the Fauna of Curagao and 224 Other Caribbean Islands 35(127):1-62. 1976. Further studies on Caribbean foraminifera. Studies on the Fauna of Curagao and Other Caribbean Islands 49(162):256 p. Hottinger, L. 1990. Significance of diversity in shallow benthic foraminifera. Atti del Quarto simposio di Ecologia e paleoecologia delle Communita Bentoniche. Museo Regionale de Scienze Naturali, Torino, 35-51. Kaehler, S., and R. G. Hughes 1992. The distributions and growth patterns of three epiphytic hydroids on the Caribbean seagrass Thalassia testudinum. Bulletin of Marine Science 51(3):329-336. Keough, M. J. 1986. The distribution of a bryozoan on seagrass blades: Settlement, growth, and mortality. Ecology 67(4):846-857. Koltes, K. H., J. J. Tschirky, and I. C. Feller In press. CARICOMP site characterization: Carrie Bow Cay, Belize, Central America. UNESCO. Langer, M. R. 1988. Recent epiphytic Foraminifera from Vulcano (Mediterranean Sea). Revue de Paleobiologie, Special Volume 2, 827-832. 1993. Epiphytic foraminifera. Marine Micropaleontology 20:235-265. EeiCalwez, J. 1936. Observations sur le genre /ridia. Archives de Zoologie Expérimentale et Génerale 78:115-131. 1938. Recherches sur les Foraminiféres. 1. Developpement et reproduction. Archives de Zoologie Experimentale et Generale 80(3):163-333. Lewis, J. B., and C. E. Hollingworth 1982. Leaf epifauna of the seagrass Thalassia testudinum. Marine Biology 71:41-49. Littler, D. S., and M. M. Littler 1997. An illustrated marine flora of the Pelican Cays, Belize. Bulletin of the Biological Society of Washington (9):1-149. Littler, D. S., M. M. Littler, K. E. Bucher, and J. N. Norris 1989. Marine Plants of the Caribbean: A Field Guide from Florida to Brazil. Washington, D.C.: Smithsonian Institution Press, 263 p. Lutze, G. F., and G. Wefer 1980. Habitat and asexual reproduction of Cyclorbiculina compressa (Orbigny), Soritidae. Journal of Foraminiferal Research 10(4):252-260. Manning, E. M. 1985. Ecology of Recent Foraminifera and ostracods of the continental shelf of southeastern Nicaragua. Unpublished Master's Thesis, Department of Geology, Louisiana State University. Martin, R. E. 1986. Habitat and distribution of the foraminifer Archaias angulatus (Fichtel and Moll) (Miliolina, Soritidae), northern Florida Keys. Journal of Foraminiferal Research 16(3):201-206. 225 Martin, R. E., and R. C. Wright 1988. Information loss in the transition from life to death assemblages of foraminifera in back reef environments, Key Largo, Florida. Journal of Paleontology 62(3):399-410. May, R. M. 1975. Patterns of species abundance and diversity. In Ecology and Evolution of Communities, edited by M. L. Cody and J. M. Diamond, 81-120. Cambridge, Mass.: Belknap Press of Harvard University. Norris, J. N. and K. E. Bucher 1982. Marine algae and seagrasses from Carrie Bow Cay, Belize. In The Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize, 1: Structure and Communities, edited by K. Riitzler and I.G. Macintyre, 167-223. Smithsonian Contributions to the Marine Sciences No. 12. Orbigny, A. D. d'. 1839. Foraminiféres. In Histoire physique, politique et naturelle de L'ile de Cuba, edited by M. R. de la Sagra, 223 p. Patriquin, D. G. 1973. Estimation of growth rate, production and age of the marine angiosperm Thalassia testudinum Konig. Caribbean Journal of Science 13(1-2):111-123. Phillips, R. C., and E. G. Mefiez 1988. Seagrasses. Smithsonian Contributions to the Marine Sciences No. 34, 104 p. Pielou, E. C. 1966. The measurement of diversity in different types of biological collections. Journal of Theoretical Biology 13:131-144. Ribes, T., and M. P. Gracia 1991. Foraminiféres des herbiers de Posidonies de la Méditerranée occidentale. Vie Milieu 41(2/3):117-126. Richardson, S. L. 1996. Epiphytic foraminiferans from Thalassia habitats, Belize, C. A. The Geological Society of America 1996 Annual Meeting, Abstracts with Programs, A-38. Sanders, H. L. 1968. Marine benthic diversity: A comparative study. American Naturalist 102(925):1-40. 1969. Benthic marine diversity and the stability-time hypothesis. In Diversity and Stability in Ecological Systems, Report of Symposium held May 26-28 1969, 71-81. Brookhaven Symposia in Biology No. 22. Schultze, M. S. 1854. Uber den Organismus der Polythalmien (Foraminiferen) nebst Bemerkungen iiber die Rhizopoden im Allgemeinen. Leipzig: Whilhelm Englemann, 1-68. Sheldon, A. L. 1969. Equitability indices: Dependence on the species count. Ecology 50(3):466-467. Short, F. T., and S. Wyllie-Escheverria 1996. Natural and human-induced disturbance of seagrasses. Environmental Conservation 23(1):17-27. Steinker, D. C., and A. L. Rayner 1981. Some habitats of nearshore foraminifera, St. Croix, U.S. Virgin Islands. Compass of Sigma Gamma Epsilon 59(11):15-26. 226 Steinker, P. J., and D. C. Steinker 1976. Shallow-water foraminifera, Jewfish Cay, Bahamas. In First International Symposium on Benthonic Foraminifer of Continental Margins, Part A: Ecology and Biology, edited by C. T. Schafer and B. R. Pelletier, 171-180. Maritime Sediments Special Publication No. 1. Tomasko, D. A., and B. E. Lapointe 1991. Productivity and biomass of Thalassia testudinum as related to water column nutrient availability and epiphyte levels: Field observations and experimental studies. Marine Ecology Progress Series 75:9-17. Tomlinson, P. B. 1974. Vegetative morphology and meristem dependence--the foundation of productivity in seagrasses. Aquaculture 4:107-130. Tomlinson, P. B., and G. A. Vargo 1966. On the morphology and anatomy of turtle grass, Thalassia testudinum (Hydrocharitaceae) I. Vegetative morphology. Bulletin of Marine Science 161(4):748-761. Vénec-Peyré, M.-T., and Y. Le Calvez 1988. Les foraminiféres épiphytes de l'herbier de Posidonies de Banyuls-sur-Mer (Méditerranée occidentale): Etude des variations spatiotemporelles du peuplement. Cahiers de Micropaléontologie, n.s., 3(2):21-40, pl. 1. Wantland, K. F. 1975. Distribution of Holocene benthonic foraminifera on the Belize Shelf. In Belize Shelf—Carbonate Sediments, Clastic Sediments, and Ecology, edited by K. F. Wantland and W. C. Pusey III, 332-399. American Association of Petroleum Geologists, Tulsa, Oklahoma, Studies in Geology No. 2. Waszczak, R. F., and D. C. Steinker 1987. Paleoenvironmental and paleoecologic implications of Recent foraminiferan distribution patterns in the lower Florida Keys. In Symposium on South Florida Geology, edited by F. Maurrass, 203-225. Miami Geological Society Memoir 3. Wilson, B. 1998. Epiphytal foraminiferal assemblages on the leaves of the seagrasses Thalassia testudinum and Syringodium filiforme. Caribbean Journal of Science 34(1-2): 131- 1132: Zieman, J.C. 1974. Methods for the study of the growth and production of turtle grass, Thalassia testudinum K6nig. Aquaculture 4:139-143. Zieman, J. C., and R. G. Wetzel 1980. Productivity in seagrasses: methods and rates. In Handbook of Seagrass Biology: An Ecosystem Perspective, edited by R. C. Phillips and C. P. McRoym, 87-116. New York: Garland STPM Press. Zohary, T., Z. Reiss, and L. Hottinger 1980. Population dynamics of Amphisorus hemprichii (Foraminifera) in the Gulf of Elat (Aqaba), Red Sea. Eclogae geologicae Helvetiae 73(3):1071-1094. APPENDIX I List of sample localities and species collected from Pelican Cays (see also Fig. 1). Water Date collected depth(m) Locality 23 August 1996 <1 Cat Cay: samples collected from off western side of the island Field # PC-96-A PC-96-B_ | 23 August 1996 0.5 Pond A: samples collected along flat projecting into bay from south, located just between Cat Cay and Cat Cay South Island 25 August 1996 0.5 Pond C (entrance): samples collected from entrance to pond PC-96-C1 PC-96-C2 | 25 August 1996 0.5 Pond C (ridge): samples collected from ridge crossing middle of pond 25 August 1996 0.5 Pond C (lagoon): samples collected from within pond PC-96-C3 PC-96-F | 27 August 1996 0.5 Pond J: samples collected on ridge at entrance to pond PC-96-G_ | 28 August 1996 <1 Pond E: samples collected at entrance to pond Type of data collected Census of epiphytic foraminiferans, shoot density, seagrass biomass, epiphyte biomass, leaf surface area Census of epiphytic foraminiferans, shoot density, seagrass biomass, epiphyte biomass, leaf surface area area Shoot density, seagrass biomass, epiphyte biomass, leaf surface area Census of epiphytic foraminiferans, leaf surface area Census of epiphytic foraminiferans, leaf surface area Shoot density, seagrass biomass, epiphyte biomass, leaf surface area, census of epiphytic foraminiferans Shoot density, leaf surface area, census of epiphytic foraminiferans 227 228 APPENDIX II List of species Belizeanella candeiana (d'Orbigny, 1839): Rosalina candeiana d'Orbigny, 1839, p. 97, pl. 8, figs. 2-4; Wantland, 1975, p. 394, figs. 10c, d, 12n; Discorbis candeiana d'Orbigny, Cebulski, 1969, p. 326, pl. 2, fig. 4. Cornuspira planorbis Schultze, 1854, p.40, pl. 2, fig. 21; Cebulski, 1969, p. 326; Wantland, 1975, p. 387. Cornuspiramia antillarum (Cushman, 1922): Nubecularia antillarum Cushman, 1922, p. 59, figs. 7, 8. Hemidiscella palabunda Bock, 1968, p. 27, pl. 4, figs. 3-9; Wantland, 1975, p. 385, figs. 10i, j. Iridia n. sp. Planorbulina acervalis H. B. Brady, 1884, v. 9, p. 657, p. 92, fig. 4; Cebulski, 1969, p. 326, pl. 2, fig.9; Wantland, 1975, p. 397, fig. 11d. Rhizonubecula n. sp. ATOLL RESEARCH BULLETIN NO. 476 DIVERSITY OF SPONGE FAUNA IN MANGROVE PONDS, PELICAN CAYS, BELIZE BY KLAUS RUTZLER, MARIA CHRISTINA DIAZ, ROB W.M. VAN SOEST, SVEN ZEA, KATHLEEN P. SMITH, BELINDA ALVAREZ, AND JANIE WULFF ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 mn 1 Hlidden Creek Guatemala a Honduras ; oA 88°W Twin Cays Be Cc es Ke| Blue Ground fone: c Sb Range ° @ ‘0: tg: _ i. Carrie Bow Cay *:: 4 : pA of vee 7 Ake of s : 0 ” a» a “th , Patch Reefs F rt & Sand Bores 0 a . a el ES es ee) AS 5km en as: z a Gio We > ine io “a she 7 eine ” Patch Reefs ™, > co & Sand Bores *. aN . es b BS One ) Figure 1. Map of Belize (a) with enlarged research areas: southern barrier reef lagoon (b), Twin Cays (c), and portion of Pelican Cays (d). DIVERSITY OF SPONGE FAUNA IN MANGROVE PONDS, PELICAN CAYS, BELIZE BY KLAUS RUTZLER,' MARIA CRISTINA DIAZ,’ ROB W. M. VAN SOEST,? SVEN ZEA,* KATHLEEN P. SMITH,' BELINDA ALVAREZ,’ and JANIE WULFF® ABSTRACT Mangrove-fringed ponds in the Pelican Cays, Belize, support an uncommonly diverse population of colorful and large sponges. Sponge species and abundance were determined for ponds at Cat Cay (Pond A), Manatee Cay (C), and Fisherman’s Cay (E and F) and compared with the sponge fauna of more typical and common mangrove habitats elsewhere in the Belize lagoon, one at Blue Ground Range and three at Twin Cays, 10—15 km to the north. The seven habitats differ in species composition and hierarchy, the Pelican Cays ponds being by far the most species-rich, with an unusually high number of poorly known or undescribed taxa. The principal factors promoting diversity in the Pelicans are abundance of solid substrates (mangrove stilt roots, extended peat banks), low turbidity, and proximity of sponge-rich coral reefs. The topography of deep ponds alternating with steep coral ridges helps contain fine sediments and prevents resuspension and silting during storms without blocking water exchange, which is necessary for importing nutrients and flushing waste. Blue Ground Range is the most sediment-exposed habitat and supports selected silting-resistant species. Twin Cays are intermediate in sedimentation and species numbers and suffer the most physical disturbance from boating and fishing. 'Department of Invertebrate Zoology, National Museum of Natural History, Smithsonian Institution, Washington, D.C. 20560-0163, USA. “Ocean Sciences Department, University of California at Santa Cruz, Santa Cruz, CA 95054, USA. ‘Institute for Systematics and Ecology, University of Amsterdam, P.O. Box 94766, 1090 GT Amsterdam, The Netherlands. ‘Universidad Nacional de Colombia (Departamento de Biologia), INVEMAR, Apartado 10-16, Santa Marta, Colombia. National Institute of Water and Atmospheric Research, Ltd., P.O. Box 14-901, Kilbirnie, New Zealand. Department of Biology, Middlebury College, Middlebury, VT 05753, USA. 232 INTRODUCTION Biologists and geologists associated with the Smithsonian National Museum of Natural History have been studying the tropical marine shallow-water communities of Belize, Central America, for more than 25 years. Most of their research has been sponsored by the Caribbean Coral Reef Ecosystems (CCRE) program and has concentrated on the barrier reef near Carrie Bow Cay (16°48'N, 88°05'W), the location of the program’s field station (Rtitzler and Macintyre, 1982), and on the mangrove ecosystem of Twin Cays (Riitzler and Feller, 1996). During the early 1990s, some program participants shifted their attention to the rich benthic communities associated with coral ridges and mangrove roots of mangrove-fringed ponds of the Pelican Cays, an archipelago of mangrove islands in the southern Belize shelf lagoon, just 16 km southwest of Carrie Bow Cay. Sponges, algae, and ascidians, are among the most abundant sessile organisms there, densely overgrowing most exposed solid substrates, and are notable for their stunning colors and bizarre shapes. All who have studied sponges in shallow waters throughout the Caribbean have been impressed by the diversity of sponges found in the Pelican Cays. To document this phenomenon, we surveyed species richness and abundance in these islands and compared them with the sponge fauna of similar environments at Twin Cays and Blue Ground Range, two groups of mangrove cays 3 km northwest and 6.5 km west, respectively, of the Carrie Bow field station and well known from previous studies. Because taxonomic work on the sponges is still in progress, we restrict our observations to the common and systematically recognized forms. A revision of the species collected will be published as part of a separate monograph. METHODS AND STUDY AREA Sponges were collected in separate bags by snorkeling and free-diving and returned to the Carrie Bow field station in buckets of fresh seawater. Where possible, specimens were photographed undisturbed in their natural setting before collecting using a Nikonos camera with close-up lens and one or two small strobe units (Ikelite) angled 45° to the camera axis. Specimens were also photographed in the laboratory submersed in a tray with seawater, with a Nikon F with Micronikkor macro lens on a copy stand and two Vivitar strobes at 45°. After a preliminary morphological study, small samples were removed for skeleton mounts and specimens were fixed in 10% formalin in seawater (12—24 h) and transferred to 80% ethylene alcohol for transport and storage. At the field station, a compound microscope was used to examine skeleton structure in dried hand sections cleared in Permount medium (Fischer Scientific). Spicule types were determined after dissolving samples in concentrated laundry bleach (5% sodium hypochlorite). Abundance estimates were based on the consensus of six collectors (Alvarez, Diaz, Van Soest, Smith, Wulff, and Zea) after each collecting trip in August 1997. The method used (1 = very common; 2 = common-rare; 3 = very rare), though subjective, captured the quantity and size of individuals a casual observer might readily notice. A small but “common” sponge (one appearing in many places) may have a similar score as a larger but less common species, whereas a single, very large specimen would still be considered a “rare” species. In comparing species composition, we applied Sorensen’s formula (Pielou, 1992) because it compensates for very different species numbers in various locations by doubling the number of shared 233 taxa: 2a/2a + b + c (where a = number of species common to both locations; b, c = number of species occurring only in one of the compared samples). Estimates of pond size were made using a planimeter on maps drawn from aerial photographs and improved and scaled by ground- truthing. Some depth measurements were made using a hand-held Scubapro PDS-2 dive sonar and converting data from feet to meters. The general setting of the Pelican Cays (16°39.8'N, 88°11.5'W) and topography and major biological components of the ponds are described by Macintyre et al. (this volume). For consistency, we use their term "pond" (designated by a letter) rather than “lagoon” to indicate the reduced water exchange with the outside by entrance-blocking coral ridges, and to avoid confusion with the principal barrier-reef lagoon. As mentioned earlier, the Twin Cays and Blue Ground Range near Carrie Bow Cay were also surveyed in this study. The mangroves, historical development, and biology of Twin Cays (16°50.0'N, 88°06.3'W) have been described by Riitzler and Feller (1987, 1996), and some work has already been done on the sponge biota, including the families Chalinidae (de Weerdt et al., 1991) and Mycalidae (Hajdu and Riitzler, 1998). The latter study includes samples from Blue Ground Range (16°48.6'N, 88°08.9'W), an extensive mangrove development at the inner (landward) edge of the barrier-reef shelf. RESULTS Four ponds enclosed or partly enclosed by red mangroves (Rhizophora mangle) were investigated at three Pelican Cays islands: Cat, Manatee, and Fisherman’s Cays (Fig. 1, Table 1). These sites were chosen because of their reportedly rich sponge fauna (I. Goodbody, personal communication) and because of time constraints. Other ponds in the Pelican group also contain sponges (Goodbody, this volume; Macintyre et al., this volume) but they are not nearly as diverse and are more similar in character to the majority of mangrove cays elsewhere in the Belize shelf lagoon. Comparative collections were made at four other sites: one at Blue Ground Range, the other three at Twin Cays. Description of Sites Cat Cay. Pond A, located at Cat Cay, is a deeply cut oval bay stretching north-south and has a shallow coral ridge and two small mangrove islets blocking most of the main entrance, which faces west and south. The coral ridge rises steeply from the sea floor and levels off at an average depth of about 0.5 m (mean tide level). The pond is about 285 x 90 m and covers more than 29,000 m? (2.9 ha). The circumference of the pond, including the coral ridge, measures more than 600 m. About 41% of the perimeter consists primarily of coral, coral rubble, and coral-covered mangrove roots; 43% is a red-mangrove shoreline with exposed and submerged roots and stilt roots; the remaining 16% is muddy intertidal or other substrate covered by algae. The coral ridge at the entrance and the coral rubble, mangrove roots, and peat banks lining the shallow margin provide a solid substrate relatively free of sediment for sponges and ascidians, the principal sessile filter feeders in Pond A. Despite the lush mangrove growth and the abundance of fine sediments, the water along the red-mangrove shoreline is extremely clear. The maximum recorded depth (at the bottom of the inner slope of the coral ridge) is 14 m; the muddy bottom near the center of the pond reaches 10.5 m. 234 Table 1. List, distribution, and estimated abundance of sponges in localities surveyed (1=rare, 2=common, 3=abundant). Taxa Homosclerophorida Plakinidae Plakina jamaicensis Lehnert & van Soest Plakinastrella onkodes Uliczka Plakortis halichondrioides? (Wilson) Oscarella sp. | Oscarella sp. 2 Oscarella sp. 3 Spirophorida Tetillidae Cinachyrella apion (Uliczka) Astrophorida Ancoriniidae Ecionemia dominicana (Pulitzer-Finali) Myriastra kallitetilla de Laubenfels Geodiiae Geodia gibberosa (Lamarck) Geodia papyracea Hechtel Geodia sp. Pachastrellidae Dercitus sp. Hadromerida Chondrillidae Chondrilla nucula Schmidt* Chondrosia collectrix (Schmidt) Clionidae Anthosigmella varians (Duch. & Mich.) Cliona caribbaea Carter Cliona sp. Placospongiidae Placospongia intermedia Sollas Spirastrellidae Diplastrella megastellata Hechtel Spirastrella coccinea (Duch. & Mich.) S. mollis Verrill Suberitidae Aaptos duchassaingi (Topsent) Terpios aurantiaca Duch. & Mich. T. fugax Duch. & Mich. T. manglaris Ritzler & Smith Terpios sp. | Terpios sp. 2 Pelican Cays Cat Noe tO NO NON — Manatee ej Se bo Lo o>) — — Fisherman's to 1S) 1vS) to i) Blue Ground Range Twin Cays a is} > vo ssi Oo age S Oo ap Ms) Ss = oS ape ao O n 2 1 2 | l 1 3 2 3 l 2 2 1 2 Table 1.--continued Taxa Suberitidae sp. Tethyiidae Tethya actinia de Laubenfels T. aff. actinia sp. | T. aff. actinia sp. 2 T. aff. actinia sp. 3 Timeiidae Paratimea? sp. Timea unistellata Topsent Agelasida Agelasiidae Agelas conifera (Schmidt) Poecilosclerida Anchinoidae Phorbas amaranthus Duch. & Mich. Coelosphaeridae Coelosphaera raphidifera Hechtel Lissodendoryx colombiensis Zea & van Soest L. aff. isodictyalis (Carter) L. sigmata (de Laubenfels) Crambeidae Monanchora arbuscula (Duch. & Mich.) Monanchora sp. Desmapsamma anchorata (Carter)* Desmacellidae Biemna caribea Pulitzer-Finali Desmacella janiae Verrill D. meliorata Wiedenmayer Desmacella sp. Neofibularia nolitangere (Duch. & Mich.) Hymedesmiidae Hymedesmia sp. Iophonidae Acarnus sp. Microcionidae Artemisina melana van Soest Clathria affinis (Topsent) C. echinata (Alcolado) C. schoenus (de Laubenfels) C. aff. schoenus (de Laubenfels) C. spinosa (Wilson) C. venosa (Alcolado) C. virgultosa (Lamarck) Clathria sp. | Clathria sp. 2 Pelican Cays 8 & iss} ) =I fe 3 Ss a Oo Ss l ] 3 l ] ] 1 3 l 1 1 2. l ] 1 A), ] 1 2 2 3 3} 1 iD) ] l 3 l 2 ] 1 l l 3 2 2) a 2 2 3 3 3 ] ] 3 3 Blue Ground Range in) NO Twin Cays a % 2 o iss} oO yy F aS ah ne} iss} i= oe 8 a0} Oo n 1 1 1 3 3 1 3 ] l l l l 2 3 235 236 Table 1.--continued Taxa Mycalidae Mycale angulosa (Duch. & Mich.) M. arenaria Hajdu & Desqueyroux-Faundez M. arndti van Soest M. citrina Hajdu & Riitzler M. escarlatei Hajdu et al. M. laevis (Carter)* M. carmigropila Hajdu & Riitzler M. laxissima (Duch. & Mich.) M. magniraphidifera van Soest M. aff. magniraphidifera van Soest M. microsigmatosa Arndt M. aff. microsigmatosa Arndt Mycale sp. | Mycale sp. Mycale sp. Mycale sp. Mycale sp. 5 Paresperella sp. Myxillidae Totrochota birotulata (Higgin) lotrochota sp. Phoriospongidae Strongylacidon sp. Raspaillidae Ectyoplasia ferox (Duch. & Mich.) Eurypon laughlini Diaz et al. Eurypon sp. Tedanidae Tedania ignis (Duch. & Mich.) T. aff. ignis (Duch. & Mich.) Halichondrida Axinellidae BW bY Pseudaxinella? spp. aff. zeai Alvarez et al. Ptilocaulis walpersi (Duch. & Mich.) Dictyonellidae Dictyonella sp. Scopalina hispida (Hechtel) S. ruetzleri (Wiedenmayer)* Scopalina? sp. Ulosa funicularis Riitzler* Halichondriidae Amorphinopsis sp.1 Amorphinopsis sp. 2 Pelican Cays 2) aa 8 &£ iss} o Ss is =} S B So fs D l 2 2) l | 2D | 3 3 3 D 1 2 3 3 ] 2 ] | | 1 ] 3 3 3 2 ] 2 1 l ] 2 2 2 2 1 2 l l 3 3 3 3 | 2) 1 2 | | Blue Ground Range Ne nN Lo Lo in) Twin Cays i=] o > o si OS 2 = os Cae Ae! Ss i=] @ Sy ee x oO n 2 1 l ] 2 | 2 | 1 2 3 I l 3 3 3 l ] l 2 3 237 Table 1.--continued Pelican Cays Twin Cays Manatee Blue Ground Range Hidden Creek Cuda Cut Sponge Haven Cat Taxa Ciocalypta? sp. Halichondria magniconulosa? Hechtel H. poa? De Laubenfels 3 Hymeniacidon caerulea Pulitzer-Finali 1 1 Myrmekioderma rea (de Laubenfels) 1 1 Topsentia ophiraphidites de Laubenfels 1 | Haplosclerida Callyspongiidae Callyspongia fallax Duch. & Mich." 2 C. vaginalis (Lamarck) 1 Callyspongia sp. l Chalinidae Halicona caerulea (Hechtel) 1 H. curacaoensis (van Soest) 1 l H. aff. curacaoensis (van Soest) 1 3 H. implexiformis (Hechtel) I H. aff. implexiformis (Hechtel) | H. magnifica de Weerdt et al. ] 1 1 H. manglaris Alcolado 2 | 2 ] 2 2 H. mucifibrosa de Weerdt et al. 2 3 H. pseudomolitba de Weerdt et al. 2 H. tubifera (George & Wilson) | 1 H. aff. tubifera (George & Wilson) H. twincayensis de Weerdt et al. Haliclona sp. Haliclona sp. 2 Haliclona sp. 3 Haliclona sp. 4 Haliclona sp. 5 Haliclona sp. 6 | 7 8 9 l l l ty —| Fisherman's ios) NO ie) N Nw Nw ww Ne WwW Se AY Haliclona sp. Haliclona sp. Haliclona sp. Haliclona sp. Haliclona sp. Haliclona sp. Chalinidae sp. | | Niphatidae Amphimedon compressa Duch. & Mich. A. erina (de Laubenfels) A. aff. erina (de Laubenfels) Gellius sp. | Niphates digitalis (Lamarck) N. erecta Duch. & Mich. Niphates sp. NO — © N WwW W wo nN Nw 238 Table 1.--continued Pelican Cays Twin Cays Blue Ground Range Manatee Fisherman's Hidden Creek Cuda Cut Sponge Haven Cat Taxa Petrosiidae Petrosia weinbergi (van Soest) Strongylophora davilai Alcolado Xestospongia carbonaria (Lamarck) X.? caycedoi Zea & van Soest X, muta (Schmidt)’ X, proxima (Duch. & Mich.) X. subtriangularis (Duchassaing) l Phloeodictyidae Aka siphona (de Laubenfels) ] Aka sp. | Calyx podatypa (de Laubenfels) Oceanapia sp. | | Oceanapia sp. 2 | Dictyoceratida Irciniidae Fasciospongia? sp. l Hyrtios proteus Duch. & Mich. Hyrtios sp. I l Ircinia campana (Lamarck) I. felix (Duch. & Mich.) 2 I. strobilina (Lamarck) l 2 l Ircinia sp. | Ircinia sp. 2 Ircinia sp. 3 | Ircinia sp. 4 | yy (Sy —_ — La — bo i) tO tO ies) NO tO Lo ies) iw) — = GW UD Smenospongia aurea (Hyatt) ] Spongiidae Cacospongia sp. | Spongia obscura Hyatt S. tubulifera Lamarck* Dendroceratida Dysideidae Dysidea. etheria de Laubenfels D. janiae (Duch. & Mich.) 3 Dysidea sp. | 3 Dysidea sp. 2 l Darwinellidae Aplysilla sp. 1 Aplysilla sp. 2 l Aplysilla sp. 3 | Chelonaplysilla atf. erecta (Row) 2 2) 3 | Darwinella rosacea Hechtel l Pleraplysilla sp. l (oS Lo NY Lo LW i) Les) Nm NO ~—) 239 Table ].--continued Pelican Cays 2) Twin Cays 5 : ee > Eee o cS 2 0 3 Bo OG oe St OS ee Sie sew Stes Taxa Sis in Se en Halisarcidae Halisarca caerulea Vacelet 2 2 2 l Halisarca sp. 2 Verongida Aplysinidae Aiolochroia crassa (Hyatt) 2 1 Aplysina archeri Higgin | A. fistularis (Pallas) 1 ] A. fulva (Pallas) 3 l 3} Verongula rigida Esper 2 2 ] Calcarea Clathrina aff. coriacea (Montagu) 2 2 3 2 Sycon sp. 1 1 ] Total number of species per locality 90 95 77 54 26 29 42 Species per region 147 54 57 *See Plate 2 Before entering the lagoon by swimming across the coral ridge, one encounters an unexpectedly steep outer slope that rises from a sandy bottom at 22 m and is covered by lettuce coral (Agaricia tenuifolia) with isolated stands of staghorn coral (Acropora cervicornis). In 1994, many of the A. cervicornis were partly bleached, possibly by white-band disease. Some isolated Agaricia blades were also freshly bleached, perhaps as a result of starfish (Oreaster) feeding, which was occasionally observed in 1997. Several large sponges are attached to the dead coral and coral rubble in this area. Clusters of large barrel sponges, dark purplish brown Xestospongia muta (Plate 2b), form several reef-like islands on the lower slope between 6 and 15 m, where coral growth is less lush and interspersed with sand, coral rubble, and patches of turtle grass (Thalassia testudinum). Associated with Xestospongia are long, intertwined ropy sponges, blackish green Jotrochota birotulata, crimson Amphimedon compressa, and ochre-yellow Aplysina fulva. Other large sponges present on the slope are tall tubes of gray Niphates digitalis and blackish red Mycale laxissima, brown bowl-shaped /rcinia campana, and massive blackish gray 1. strobilina. In the shallow parts, and on top of the Agaricia ridge, coral is partly overgrown by extensive patches of greenish brown Chondrilla nucula (Plate 2a), a leathery and tough encrusting sponge with a very smooth, slick surface, which is known to be an aggressive competitor for space (Vicente, 1990). (This sponge may well be an undescribed species of Chondrilla because its growth habit and ecology are very different from that of the familiar C. nucula, a Mediterranean relative, although the siliceous spicules, spherasters, are indistinguishable between the two.) Another green encrusting sponge in this habitat, Ulosa 240 funicularis (Plate 2a), is darker (with a grayish tinge) and soft, and has thin whip-like processes (Riitzler, 1981). Both harbor large numbers of unicellular cyanobacteria and thus are photosynthetic species (Riitzler, 1990). These two crustose species—along with ropy, yellow Aplysina fulva and blackish Jotrochota birotulata, tube clusters of violet to pink Callyspongia fallax, and layers of a colonial gray-green zoanthid, Zoanthus sp.—also grow on loose coral plates and seem to help stabilize them. Riitzler (1965) introduced the term “Kittschwaémme” (“putty” sponges) to describe this sponge-mediated binding of reef coral and rubble, a phenomenon also described by Wulff and Buss (1979) and Wulff (1984). Other sponges attached to Agaricia coral plates on the ridge are encrusting red Monanchora arbuscula and cushion- shaped yellow Mycale laevis (Plate 2c,d), as on open reefs, attached to the underside of plates, and blackish-olive Xestospongia carbonaria. At least one excavating sponge, the dark brown symbiotic (with zooxanthellae) Cliona caribbaea is commonly found in the coral skeleton. Moving clockwise from the ridge toward the pond, one finds red-mangrove root and rubble substrates overgrown by coral (forming Agaricia “mini-reefs”) and by large or extensively encrusting sponges, particularly branching pink Desmapsamma anchorata (Plate 2c), blackish green (exuding purple stain when touched) Jofrochota birotulata, many covered by a yellow- orange symbiotic zoanthid (Parazoanthus swifti), deep orange-brown Artemisina melana, deep red Mycale laxissima, blackish Xestospongia carbonaria, and bright crimson Amphimedon compressa. Massive or thickly encrusting forms include the ubiquitous orange-red Scopalina ruetzleri, blackish-with-gray large (80 cm across) /rcinia strobilina, dark wine-red Mycale laxissima, sky-blue Dysidea etheria, and yoke-yellow Mycale laevis (Plate 2c,d). The most common thinly encrusting species are blue-green Terpios manglaris, grayish to pinkish Clathria echinata, and red Monanchora arbuscula. Massive sponges such as brown /rcinia felix and black Hyrtios proteus are often encountered on coral rubble between the mangrove fringe and stands of turtle grass. Excavating Cliona caribbaea occur in the same habitat. Continuing clockwise around the pond, one finds at least 90 species, but no distinctive distributional patterns can be discerned. The species just mentioned occur here as well. The larger massive forms include the amorphous black Spongia tubulifera, which supports a great variety of smaller epizoic sponges, bluish-green Amphimedon erina, sprawling-tubular brown Calyx podatypa, yellow Aiolochroia crassa and Verongula rigida, branching reddish black Artemisina melana, and yellow-over-salmon Clathria schoenus. Conspicuous crusts on mangrove roots are pale-red Spirastrella mollis, deep red Phorbas amaranthus, red-granular Mycale microsigmatosa, and leathery blue Halisarca caerulea. Manatee Cay. The sponge survey at Manatee Cay focused exclusively on the large lagoon-like “Pond C,” which is actually a composite of three bays that may have evolved from separate ponds. The pond measures 333 m (N-S) x 144 m (W-E) and covers an area of 3.24 ha. Most of the 638-m long circumference is covered by a mangrove fringe. The 60-m wide entrance, which faces west, is blocked by the shallows of a coral ridge. The largest, northern lobe of the pond (which we have named C, and covers 1.87 ha, with 485 m of shoreline) is the richest in terms of habitat variety and populations of filter-feeding sponges and ascidians. The entrance into C, faces south and is narrower than the main entrance (44 m). This pond is deeper and visibility is lower than in Cat Cay's pond A. During our survey (August 5, 1997) we measured the following maximum depths (by hand-held sonar) and average horizontal visibility at the surface (Secchi disk): 241 Just west outside main entrance into Pond C 15.0 m depth/8.0 m visibility Center of northern lobe (C,) 11.0 m depth/4.5 m visibility Center of Pond C (east of main entrance) 10.0 m depth/6.0 m visibility Just east of center of southern lobe 13.5 m depth/4.5 m visibility Turning north and swimming clockwise after entering Pond C, one first encounters a muddy bottom with turtle grass and Halimeda algae. Sponges are associated with pieces of rubble and include brown clumps of Anthosigmella varians, greenish brown crusts of Chondrilla nucula (Plate 2a), clusters of dull green Amphimedon erina and brownish Ircinia felix, and red-orange, cake-shaped Lissodendoryx colombiensis. Red mangrove roots in this area support primarily the red encrusting Mycale microsigmatosa, branching red-blackish Artemisina melana, and blackish Totrochota birotulata. From here on, most of the western submerged shoreline consists of more or less undercut peat banks with exposed mangrove roots and overhanging or overarching Rhizophora stilt roots. Both are densely carpeted by sponges and ascidians; sponges are dominant and exhibit considerable diversity in the darker parts of the undercuts where there is no competition for space from algae. Common exposed massive species, apart from Artemisina and Jotrochota, are yellow Aplysina fulva, black Spongia tubulifera (Plate 2d) and Hyrtios proteus, yellow Mycale laevis (Plate 2c,d), some surviving half-buried in mud after having fallen off the original substrate, orange Scopalina ruetzleri (Plate 2e), red Clathria schoenus, red fire sponge Tedania ignis, and dark-brown Xestospongia proxima. Large, continuous but very thin crusts include pale red Spirastrella mollis, blue-green Terpios manglaris, and red Monanchora arbuscula. Smaller but diverse and abundant encrusting and cushion-shaped species occur as epizoans on other species ( e.g., on Spongia tubulifera) and particularly on the back walls and ceilings of the cave-like undercuts: brown Placospongia intermedia, grayish Clathria venosa, greenish and reddish species of Clathria and Mycale, orange Scopalina ruetzleri (Plate 2e). Again, crusts are common in the dark zone of the caves: gray to whitish Haliclona curacaoensis, bluish-green Amphimedon erina, blackish olive Xestospongia carbonaria, blue Halisarca caerulea, and near-spherical, green and orange Tethya actinia. Many of the large and common species extend their distribution to the northeastern, southeastern, and southern flanks of the pond where habitats are restricted to stilt roots isolated by mud bottoms, but diversity is markedly lower then in the zone of the peat undercuts. Only a few sponge species, such as the common Scopalina, Ircinia, and Tedania, occur along the central eastern shoreline across from the main entrance. Fisherman’s Cay. Two interconnected ponds, E and F, are present here, but only E, the northern one, has an entrance to the open water. The two ponds are almost the same size and together form a figure 8: E with an area of 0.40 ha (225-m circumference), F covering 0.41 ha (227-m circumference). The longest perpendicular axis of E measures 122 x 44 m and that of F is 91 x 70 m. The only entrance from the outside is 8 m wide and up to 4 m deep; the center of pond E drops to 11 m. The Agaricia coral ridge has a steep slope with some conspicuous clusters of tubes of the blackish red sponge Mycale laxissima. There is a forest of stilt roots at the entrance and ample coarse sand made up of Halimeda chips. Some turtle grass, Thalassia, is also present, and although the slope into the pond is muddy, the water is surprisingly clear. Roots and peat banks with small undercuts at the entrance are thickly covered by algae, ascidians, and large encrusting or ropy branching sponges, particularly the orange Scopalina ruetzleri (Plate 2e), grayish Clathria 242 venosa, yellow Aplysina fulva, and blackish Artemisina melana. Hanging mangrove roots backed by a vertical peat bank 1—2 m in relief, line most of the pond and provide substrate for the sessile organisms, including algae, ascidians, sponges, corals, and some octocorals. Prominent sponges are massive yellow Mycale laevis (Plate 2c,d), blackish Spongia tubulifera and S. obscura, ropy red Clathria schoenus, and encrusting brown Chondrilla nucula (Plate 2a) and red Mycale microsigmatosa. The passage connecting E and F is only 5 m wide and less than 1 m deep. Again, the roots are covered thickly by S. ruetzleri (Plate 2e) and draped by A. fulva, and coral rubble at the bottom contains a yellow excavating sponge, Cliona sp. Other conspicuous sponges found here are thinly encrusting grayish-yellowish Clathria venosa, several Mycale spp. (including red M. microsigmatosa), blue Halisarca caerulea, and a second tan species of Halisarca. Many cushion-shaped or thickly encrusting smaller species are also present, including deep purple Chelonaplysilla cf. erecta, blue and gray Dysidea etheria and Dysidea sp., blue Haliclona curacaoensis, blackish green Jotrochota birotulata and Xestospongia carbonaria, gray-green Lissodendoryx cf. isodictyalis, and tan, ball-shaped Cinachyrella apion. Other Sites. Using the same techniques, we sampled four sites in two other mangroves, Blue Ground Range and Twin Cays. Both ranges were familiar to some of us from previous studies. They have an abundance of habitats supporting a diverse sponge fauna, and seem more typical of the thousands of mangrove cay habitats scattered throughout the Belize lagoon than in the Pelicans. The primary difference is that most of the mangrove islands do not have the deep, clear, isolated ponds characteristic of the Pelican Cays but instead have a system of tidal channels and shallow lakes with high turbidity from fine sediments that are suspended and redeposited with each tidal cycle and storm event. The sponges sampled are listed in Table 1. Blue Ground Range (locally also known as Cockney Range) is a series of mangrove islands lying in a line oriented north to south. Our site (“Center of Origin”; Hajdu and Riitzler, 1998) is an extensive, deep (>5 m) lagoon in the center of the range that can be entered from the east by a large boat but has only a very shallow (0.5-m) cut facing west. The cut can only be crossed by swimming but allows good water exchange when the tide changes and the wind blows from the east or west. The bottom is composed of a thick layer of burrowed mud and the water is usually turbid. Sponge habitats are restricted to hanging and anchored stilt roots from a red mangrove and peat bank along the western flank of the lagoon, on either side of the shallow westward cut. At Twin Cays, northwest of Carrie Bow Cay (Riitzler and Feller, 1987, 1996; de Weerdt et al., 1991; Hajdu and Riitzler, 1998) we sampled three sites. One, Hidden Creek, is a 2-m deep, meandering tidal creek that connects an open bay with a shallow enclosed mangrove lake. During each tidal cycle, the water flowing through is either hot and saline (when low tide occurs during a sunny day) or cold and brackish (when low tide occurs during a rainy cold spell) and is exchanged for more tempered open-lagoon water. Values for temperature, salinity, suspended sediments, water flow, and dissolved organics (tannins) can be extreme. Hidden Creek is lined by red mangrove stilt roots and vertical or undercut peat banks covered by sponges. The second site, Cuda Cut, is a short but wide and relatively deep (4 m) passage between the open lagoon and the main channel that separates the two islands that give Twin Cays their name. Water is cool and relatively clear because it is regularly exchanged by tidal and wind currents, with modest sedimentation despite a heavily burrowed, fine-muddy bottom. There are the usual red-mangrove stilt roots and, on one flank, a deeply undercut peat bank with exposed mangrove roots, all 243 covered by sponges. The third site, Sponge Haven, is a shallow (1-m) bay off the main channel and close to its southern exit. Its conditions are intermediate; it is less exposed to wind and is flushed by fresh seawater less often than Cuda Cut. Furthermore, it is not subject to regular environmental extremes, as Hidden Creek is. There are ample mangrove roots and a low-relief peat bank and sponges are plentiful, as the name implies. Sponge Richness and Abundance Table 1 provides a preliminary list of the taxa collected and now under study. The list and distributional data are the result of our group survey in August 1997 and are supplemented by a few records obtained during earlier visits (by K. Riitzler in 1994, and M. C. Diaz and K. P. Smith in 1996, both unpublished). At this stage, we count 182 species and forms for all seven surveyed localities combined, although only 100 (55%) are readily identifiable. The remaining 45% of the sponges are either new species or morphological variants caused by adaptations to the unusual physical and chemical environment in mangrove ponds or tidal canals. Our sampling locations are comparable in size (each can be surveyed by snorkeling in about two hours) and they appear similar in the biomass of sponges growing on mangrove roots and peat banks. However, they clearly differ in species composition and richness. The Pelican Cays have the highest number of species combined (147 species) as well as per island and pond studied; Manatee Cay has the highest number (95); Cat and Fisherman’s cays are second (90) and third (77). Blue Ground Range is next in the hierarchy (54), being richer than each of the Twin Cays sites but not richer than all combined (57). Among the Twin Cays locations, Sponge Haven has the most species (42), Cuda Cut is next (29), and Hidden Creek has the fewest (26). These mangrove islands provide habitats for 30 very common and quantitatively important sponge species, but the distribution of most is not uniform. Only four species are dominant in all three regions (Pelican, Blue Ground, and Twin): Tedania ignis, Clathria venosa, Scopalina ruetzleri (Plate 2e), and Hyrtios proteus. Shared exclusively by Pelicans and Blue Ground are three abundant species, Lissodendoryx colombiensis, Artemisina melana, and Mycale arenaria. Common to both Pelican and Twin cays but rare or not recorded at Blue Ground are Plakortis halichondrioides?, Spirastrella mollis, Tethya actinia, Haliclona curacaoensis, and Amphimedon erina. Abundant in the Pelicans but lacking or very rare elsewhere are 11 species: Placospongia intermedia, Monanchora arbuscula, Desmapsamma anchorata (Plate 2c), Scopalina hispida, Topsentia ophiraphidites, Callyspongia fallax (Plate 2f), Xestospongia carbonaria, X. proxima, Aiolochroia crassa, Aplysina fulva, and Verongula rigida. Exclusive to Blue Ground are four common species: Acarnus sp., Eurypon laughlini, Ircinia campana, and Dysidea janiae. And important at Twin Cays but uncommon or missing elsewhere are Geodia papyracea, Biemna caribea, Mycale aff. magniraphidifera, and Halichondria magniconulosa. We paired the seven sampling localities and three island regions and compared them with each other using Sorensen’s similarity coefficient (Pielou, 1992) (Table 2). The three islands of the Pelican group agree well with each other (between 56 and 61%), as do the three locations within Twin Cays (55-56%). Blue Ground Range does not agree well with the other locations but it is closer to Twin Cays (45%) than to the Pelicans (35%). Comparing Blue Ground with individual sites, Manatee among the Pelicans and Sponge Haven of Twin Cays share the most species (46% and 42%, respectively). The Pelican and Twin Cays (all sites combined) share only 244 Table 2. Similarity matrix (Sorensen’s Coefficient) showing agreement of sponge species (%) collected at all seven localities (and three regions) surveyed. Shadings (dark to none) indicate <50%, 40-50%, 30-40%, 230% (— = Comparison not applicable.) Pelican Cays ' Twin Cays 8p 5 x “ 5 Ses < 3 a Rees | F Bs 5 ite) 6p oO Zz op rs) » = 5 5 E 3 3 =I S E Localities 5 Si fe 5 a = 5 & 5 Cat — 60.5 56.3 36.1 19.0 235) 30.3 37) Ss 5 Manatee 33.6 36.8 c & Fisherman’s 38.7 47.8 oO a Combined 27.5 28.4 Blue Ground Range 41.7 45.1 Hidden Creek 55.9 Cuda Cut 56.3 — Sponge Haven Twin Cays Combined 28% of the species. Broken up into individual locations, Fisherman’s Cay has the closest ties to Twin Cays (48%); Cat Cay agrees the least (33%). DISCUSSION The primary purpose of our present survey was to determine species richness and frequency of occurrence of sponges in the studied locations. According to our estimates, the Pelican Cays harbor the highest concentration of sponge species and biomass per unit area known to us in the entire Caribbean. At this point we can only speculate on the reasons for this phenomenon. One plausible explanation is that the Pelicans are close to well-developed coral reefs, a reservoir of a diverse sponge fauna that is not as readily available at the other sites because sponge larvae are unable to cross long stretches of uninhabitable environment. Indeed, several of the sponges flourishing in the Pelican ponds—such as Callyspongia vaginalis, Amphimedon compressa, species of Ectyoplasia, Topsentia, Monanchora, Myrmekioderma, Niphates, Xestospongia, and representatives of the Aplysinidae—are typical of those on nearby reefs. However, even if larvae reach a new habitat they will only settle and survive if circumstances are suitable. Our observations and preliminary data from a previous study (Diaz and Smith, unpublished report 1996) help to clarify some essential environmental differences 245 among the surveyed sites. In order to thrive, sponges require solid substrates for settlement and growth, low sediment exposure to avoid clogging of ostia, modest water movement to prevent silting or physical damage and to provide for food and flushing of waste, and little pressure from space competitors and predators. Primary substrates in these habitats are mangrove roots, mainly Rhizophora stilt roots that are either hanging free or anchored in the bottom, and peat banks, a conglomerate of roots, hair rootlets, detritus, sand, and mud exposed through erosion and in places undercut by currents to form cave-like habitats. Stilt roots are particularly suitable substrates for sponges because they allow settlement over a broad area (offering different light intensities) and depth range (from low- tide level to areas near the silty bottom). Some specialized species with a modest tide range (50 cm) are able to pioneer into the intertidal zone in these mangroves and live and reproduce where sponges are rarely found (Riitzler, 1995). Counts of roots (Diaz and Smith, unpublished) reaching below the water surface along the mangrove fringe of the ponds average from 2.2 roots per linear meter at Twin Cays (Sponge Haven) to 3.6 roots/m at Blue Ground Range and 4.2 to 6.2 roots/m in the Pelican Cays. Peat banks too are most common and best developed in the Pelican Cays ponds. These solid substrates combined with low turbidity account in large part for the greater species richness in the Pelican Cays, although Diaz and Smith report that on a species per root basis, the differences are not pronounced: 1.6 to 2.2 species/root at the Pelicans, versus 1.4 species at Blue Ground, and 2.0 species at Twin Cays. Blue Ground Range, although rich in suitable substrate, is adversely affected by high turbidity and abundant fine sediments that are readily resuspended by storms and other disturbances. At the same time, Blue Ground is rich in sponge biomass and harbors an unusual spectrum of species that seem resistant to the effects of sedimentation and capable of growing to considerable size. With less substrate and moderate turbidity, Twin Cays exhibits comparatively low sponge diversity. Even so, the average number of individuals per root is highest at Twin Cays (4.2, versus 1.9 to 3.4 at Pelicans and 1.6 at Blue Ground), perhaps because of high population turnover related to frequent disturbance by boats and dragging of seines in these heavily fished cays. The surprisingly low turbidity in the Pelican Cays (for Belize- lagoon environments) may be due to the bottom topography of deep ponds and steep honeycomb- like coral ridges, which prevent excessive resuspension of fine sediments during storms and other disturbances. Of the Pelican Group, Fisherman’s Cay is closest to Twin Cays in turbidity and species overlap (47.8%), probably because its ponds (E and F) are small and enclosed, less flushed by open-lagoon waters, and more enriched by organic compounds released from the mangroves. Despite the clarity of the water, there is sufficient food to support dense populations of filter feeders (sponges, ascidians, bivalves, in particular), which depend largely on plankton organisms, bacterioplankton, detritus, and possibly dissolved organics. In addition, sponges probably make efficient use of their resources, through their bimodal pattern of particle retention, which enables them to capture bacterioplankton (0.3 to 1.0 um) with their choanocytes as well as larger particulate organics (up to 50 um; eukaryotic cells, microscopically unresolvable organic particles) by phagocytosis in the inhalant canals (Reiswig, 1971). Competition for space appears similar in all study locations and does not explain sponge- faunal differences between sites. Sponges compete quite successfully in this regard, and many species tolerate overgrowth by other sponges (Riitzler, 1970) and by other sessile organisms such as algae, hydroids, and ascidians. Despite the space crunch in some locations, a surprising amount 246 of substrate recently formed or exposed remains unoccupied or undiscovered, perhaps because, as Zea (1993) determined for sponges in Colombian rock and reef habitats, larvae are not constantly released, are weak swimmers, and have a short planktonic life. As recently suggested by Maldonado and Uriz (1999), dispersal by means of embryo-carrying fragments is not an effective alternative in these stagnant pond waters, although fragmentation of embryo-rich sponges does occur in a few species (e.g., Scopalina ruetzleri, Plate 2e) under adverse environmental conditions such as temperature stress. This poor dispersal and recruitment ability may also account for local per-root, within-pond, or inside-range variation in sponge composition and abundance (Zea, 1996). Algae, a highly diverse group in these islands (Littler et al., this volume), are strong competitors in many of the habitats studied, particularly since they are not as heavily grazed as on reefs, although the low light levels among the roots and under peat banks slow their growth. Predation pressure, on the other hand, has only recently been recognized as an important determinant of sponge distribution (Wulff, this volume). The variable ecological setting in near- mangrove environments may influence the abundance of sponge predators such as asteroid echinoderms and angel fish and thus help to explain important differences in the sponge fauna of different mangrove cays. CONCLUSIONS The numerous mangrove cays and ranges in the Belize shelf lagoon provide important habitats for marine shallow-water communities. Sponges are a significant part of the fauna: they are rich in species and biomass; provide substrate, shelter, and defense for many other organisms; and affect habitat structure and environmental quality through effective space competition and filter-feeding. The Pelican Cays harbor the most diverse sponge fauna of all the mangrove islands previously visited or studied because of their proximity to the cache of species in nearby reefs. Furthermore, their deep mangrove-fringed ponds provide ample solid substrates, low exposure to sedimentation, and high levels of microplankton suitable for filter feeders. The special topography of steep coral ridges and deep mangrove ponds helps to stabilize sediments in these habitats, which are rich in fine detritus and carbonate mud and sand, even during storms. However, protection from frequent and careless activities by boating, fishing, and snorkeling will be important to preserve the diversity, abundance, and health of these ecologically fragile and delicately balanced communities. ACKNOWLEDGMENTS We thank Paul and Mary Shave, Ivan Goodbody, and Mike Carpenter for sharing their knowledge of the Pelican Cays ponds and Larry Manes for acting as station manager and boat captain during the period of our survey. Tony Rath and Jimmy Smith provided invaluable aerial photography of the Pelican Cays. We are grateful to Molly Ryan for drawing navigation maps and Figure 1. British Airways and Trans World Airlines made substantial contributions by donating airline tickets for Belinda Alvarez and Rob van Soest, and for Maria Cristina Dias and Larry Manes, respectively. Sven Zea’s work is Contribution No. 157 of the Marine Biology Graduate Program of the Universidad Nacional de Colombia, Faculty of Sciences; and No. 626 of the Instituto de Investigaciones Marinas y Costeras “José Benito Vives de Andreis’>-INVEMAR. 247 Contribution No. 586, Caribbean Coral Reef Ecosystems Program, National Museum of Natural History, Smithsonian Institution. REFERENCES Hajdu, E., and K. Riitzler 1998. Sponges, genus Mycale (Poecilosclerida: Demospongiae: Porifera), from a Caribbean mangrove and comments on subgeneric classification. Proceedings of the Biological Society of Washington 111:737-773. Maldonado, M., and M. J. Uriz 1999. Sexual propagation by sponge fragments. Nature 398:476. Pielou, E. C. 1992. Biogeography. Malabar, Fla.: Krieger, 351 pp. Reiswig, H. 1971. Particle retention in natural populations of three marine demosponges. Biological Bulletin 141:568—-591. Riitzler, K. 1965. Substratstabilitat als 6kologischer Faktor im marinen Benthos, dargestellt am Beispiel adriatischer Poriferen. /nternationale Revue der gesamten Hydrobiologie 50:281—292. 1970. Spatial competition among Porifera: Solution by epizoism. Oecologia (Berlin) 5:85—95. 1981. An unusual blue-green alga symbiotic with two species of Ulosa (Porifera: Hymeniacidonidae) from Carrie Bow Cay, Belize. Marine Ecology 2:35—50. 1990. Associations between Caribbean sponges and photosynthetic organisms. In New Perspectives in Sponge Biology, edited by K. Riitzler, 455-466. Washington, D.C.: Smithsonian Institution Press. 1995. Low-tide exposure of sponges in a Caribbean Mangrove Community. P.S.Z.N. I: Marine Ecology 16: 165-179. Riitzler, K., and I. C. Feller 1987. Mangrove swamp communities. Oceanus (Wood Hole) 30(4):16—24. 1996. Caribbean mangrove swamps. Scientific American, March, 94-99. Ritzler, K., and I. G. Macintyre (eds.) 1982. Habitat distribution and community structure of the barrier reef complex near Carrie Bow Cay. The Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize 1: Structure and Communities, edited by K. Riitzler and I. G. Macintyre, 9-43. Smithsonian Contributions to the Marine Sciences, v. 12. Wicente: V. P. 1990. Overgrowth activity by the encrusting sponge Chondrilla nucula on a coral reef in Puerto Rico. In New Perspectives in Sponge Biology, edited by K. Riitzler, 436-442. Washington, D.C.: Smithsonian Institution Press. Weerdt, W. de, K. Riitzler, and K. P. Smith 1991. The Chalinidae (Porifera) of Twin Cays, Belize, and adjacent waters. Proceedings of the Biological Society of Washington 104:189-205. Wulff, J. L. 1984. Sponge-mediated coral reef growth and rejuvenation. Coral Reefs 3:157-163. 248 Wulff, J. L., and L. W. Buss 1979 Do sponges hold coral reefs together? Nature 281:474-475. Zea, S. 1993. Recruitment of demosponges (Porifera, Demospongiae) in rocky and coral reef habitats of Santa Marta, Colombian Caribbean. Marine Ecology 14: 1-21. 1996. Random patterns of sponge distribution in remote, oceanic reef complexes of the Southwestern Caribbean. Abstracts, 8" International Coral Reef Symposium, Panama, June 24-29, 1996, p. 215. ATOLL RESEARCH BULLETIN NO. 477 SPONGE PREDATORS MAY DETERMINE DIFFERENCES IN SPONGE FAUNA BETWEEN TWO SETS OF MANGROVE CAYS, BELIZE BARRIER REEF BY JANIE L. WULFF ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 Mexico ms y G Caribbean & a A Sea 17 N ’ y Hlidden Creek Guatemala a Honduras ies 29° 5 | : O15 ArT Twin Cays os re Blue Ground = : SS Range * ¢ 20: arrie Bow Cay: JE, B. ‘ . q 6 re Diekne —~ a oot 0 ae de . SoyRe Patch Reefs * & Sand Bores i : 5 km adie tsi: 7 Te eee Cid, | a GIN 5 Ses Oe . a is; Pie = 7 F ” Patch Reefs, & Sand Bores *. b aos Ds Gliseaaaioroarn Be: = Figure 1. Map of Belize (a) with enlarged research areas: southern barrier reef lagoon (b), Twin Cays (c), and portion of Pelican Cays (d). SPONGE PREDATORS MAY DETERMINE DIFFERENCES IN SPONGE FAUNA BETWEEN TWO SETS OF MANGROVE CAYS, BELIZE BARRIER REEF BY JANIE L. WULFF' ABSTRACT Mangrove roots in the well-protected channels and ponds among dense mangrove stands provide very similar habitats for sponges in the Twin Cays and Pelican Cays and harbor extraordinarily dense and diverse sponge communities in both locations. The species of sponges are very different, however, possibly because of an important difference in context: the Twin Cays mangrove roots are embedded in peat banks, whereas the Pelican Cays mangrove roots are embedded in coral reefs, which may provide habitat for sponge-feeding fishes. Current and previous observations of feeding preferences and habitat requirements of potential sponge predators suggest that sponge predators play a key role in determining substantial differences in the sponge faunas, both in species composition and in species diversity, of Twin Cays and Pelican Cays. INTRODUCTION Mangrove cays on the Belize Barrier Reef appear to be close to an ideal habitat for many filter-feeding animals, among them an extraordinary diversity and density of sponges (Riitzler and Feller, 1996). Physical and chemical factors are highly favorable, with the mangroves providing protection from physical disturbances, such as storm waves, and also providing an organically enriched environment favoring organisms that filter their food out of the water column. These factors are important determinants of the distribution and abundance patterns of tropical sponges. For example, areas that are relatively protected from physical disturbance, either geographically (i.e., out of the hurricane belt) or in lagoons and on leeward sides of islands, are the only areas in the Caribbean in which significant sponge abundance has been reported in very shallow (< 2-m depth) water (e.g., Alcolado, 1979; Alvarez et al., 1990; Wulff, 1995a, and additional references therein). Nutrient availability has also been related to sponge abundance on geographic spatial scales (e.g., Wilkinson and Evans, 1989; Zea, 1994). Differences in sponge distribution among adjacent habitats with similar nutrient levels and protection from adverse physical factors have been recently shown to depend on the interactions of sponges with other organisms, especially predators. For example, the community of sponges found in seagrass meadows and rubble beds inhabited by the large starfish Oreaster reticulatus (Fig. 2) is limited to those sponge species with inherent defenses against predation by this starfish (Wulff, 1995b); and the starfishes are efficient at finding and consuming undefended sponge species that stray into their habitat from adjacent reefs. Similarly, some species of cryptic ‘Biology Department, Middlebury College, Middlebury, VT 05753, USA. Figure 3. Gray angelfish, Pomacanthus arcuatus, looking for prey (photo: Carl Hansen). 253, sponges inhabit crevices in the reef because they are refuges from the attentions of generalist herbivores, such as parrotfish of the genus Sparisoma in the Caribbean (Wulff, 1997a) and of the genus Scarus, and from omnivores such as large smooth pufferfish of the genus Arothron in the tropical eastern Pacific (Wulff, 1997b). A very different relationship exists between exposed reef sponges and spongivorous fishes. A large percentage of exposed reef sponge species is consumed by angelfishes of the genera Holacanthus and Pomacanthus (Fig. 3), the most important sponge specialist fishes in the Caribbean (Randall and Hartman, 1968; Wulff, 1994). Unmanipulated angelfishes in a natural reef sponge community were observed to feed on 36 of the 42 species present (Wulff, 1994). Exposed reef sponges continue to coexist with the angelfishes because the fishes take only small amounts of each species, possibly to avoid being poisoned by defensive chemistry (Randall and Hartman, 1968; Wulff, 1994). Even though the angelfishes do not consume reef sponges to the point of eliminating them, they do exhibit preferences such that the rate at which sponge species are consumed is inverse to their abundance (Wulff, 1994). This suggests that the angelfishes may exert some control over the relative abundance of sponge species. That control would be expected to be extreme in the case of sponge species that have not been selected to develop and maintain inherent defenses because they do not live in habitats with specialist sponge predators. This expectation appears to be corroborated by the findings of Dunlap and Pawlik (1996), who made four species of mangrove sponge species available to angelfishes, which rapidly consumed one of them. The reputation of sponges as inedible therefore appears to derive from a consistent pattern of being adequately defended against predators with which they coexist. Membership in sponge faunas typical of habitats such as coral reef, seagrass meadows, rubble beds, sediment plains, and mangrove roots, may at least in part be determined by which predators have access to these habitats. Any departure from the normal complement of sponge species in a habitat may therefore reflect unusual circumstances with respect to predator access. In the Pelican Cays, the mangroves are rooted in coral reefs instead of in peat banks (as in the Twin Cays) or in terrestrial sediments (as on the mainland). In the complex microtopography provided by the corals, sponge-feeding fishes may find shelter from their predators and thus gain access to mangrove root sponge assemblages that are normally off limits because of risk of predation. METHODS During late summer of 1997, intensive collecting activity focused on three sites in the Twin Cays (Hidden Creek, Cuda Cut, and Sponge Haven), as well as three sites in the Pelican Cays (Cat Cay, Manatee Cay, and Fisherman’s Cay). The field team was made up of Caribbean sponge biologists Belinda Alvarez, Cristina Diaz, Rob van Soest, Kate Smith, Janie Wulff, and Sven Zea. All sponge species observed were listed, collected, and their relative abundance noted (see Riitzler et al., this volume). In addition, I recorded microhabitat distributions, associations, evidence of predation, and presence of potential predators. Direct evidence of predation consisted of actual feeding on sponges and, also in the case of starfishes, the typical rounded feeding scars left by the extruded stomachs of Oreaster reticulatus (Wulff, 1995b). The potential for predation on sponges was assessed from qualitative surveys of the relative abundances of known sponge predators, including starfishes, parrotfishes, spadefishes, angelfishes, trunkfishes, and filefishes. These include all Caribbean species that have been observed to feed on sponges (Wulff, 1994, 1995b, 1997b) or to have sponge remains in their gut contents (Randall and Hartman, 1968). 254 RESULTS Sponges Sponges were qualitatively ranked according to increasing relative abundance (see Riitzler et al., this volume) on a scale of 1 to 3, by consensus of the field team after each field trip. Of the 23 species judged to be the most common sponges at Twin Cays (Table 1), 6 were even more abundant in the Pelican Cays, and the other 17 species were as common or more common at Twin Cays than in the Pelicans. Of those 17 species, 5 (Biemna caribea, Mycale aff. magniraphidifera, Halichondria ?poa, Haliclona pseudomolitba, and Haliclona mucifibrosa) were not reported at all from the Pelicans. By contrast, of the 30 most common sponge species in the Pelicans, only 3 were also common at Twin Cays, and 16 were not reported at all from Twin Cays (Table 2). The majority of the species on mangrove roots in the Pelicans were not recorded at Twin Cays. This reflects a dramatic difference in overall diversity, with 2.6 times as many sponge species found in the Pelicans (147 species and distinct forms recorded in the Pelicans, versus 57 species and forms at Twin Cays; Table 2). Potential Sponge Predators Although potential sponge predators were observed at five of the six sites, their abundance, and the degree to which different groups were represented, differed (Table 3). No individuals of fish groups that are known to specialize on sponges (i.e., angelfishes, trunkfishes, filefishes, and spadefishes) were observed in the Twin Cays; by contrast, large individuals representing all of these groups were observed in the Pelicans. Especially striking in Fisherman’s Cay Pond E were many large grey angelfishes, Pomacanthus arcuatus, and a few enormous French angelfishes, P. paru, as well as two large spadefishes, Chaetodipterus faber. In Manatee and Cat Cays, large angelfishes were positioned at intervals under overhanging peat banks. No sponge specialist species were observed at Twin Cays. In Twin Cays, small parrotfishes of the species Scarus isertii and Sparisoma radians were seen among seagrass at Cuda Cut, but heavy epiphytization on the especially dense Thalassia testudinum blades may provide evidence of their relative scarcity. Medium-large individuals of several species of the parrotfishes genus Sparisoma were seen in all of the ponds in the Pelicans, but the parrotfishes were more common on the reefs. The large starfish Oreaster reticulatus was found in Twin Cays, in areas of less dense seagrass and occasionally on the peat banks into which mangrove roots were embedded. However, this starfish was far more abundant in the Pelicans, where individuals are especially concentrated directly under the mangrove roots in areas with patches of hard substrate or in which sediment is stabilized by seagrasses. Observed Predation on Sponges Individual Oreaster reticulatus were observed to be feeding on, or had left unambiguous scars from recent feeding (nearly round areas, 8—16 cm in diameter, from which live tissue had been recently removed) on 11 species of sponges (Table 4). In several cases, sponges appeared to have very recently fallen off mangrove roots and up to 3 starfishes were observed to be feeding on one of these large windfalls at a time. In one case, a chunk of the highly edible (Wulff, 1995b) Table 1. The most abundant 23 sponge species on mangrove roots in Twin Cays, Belize, and all species reported from mangroves in at least 3 other Caribbean locations. Locations Bah Jam Cub VeM VeB PaS_ PaG Sponge species | Bel Plakortis halichondrioides? Xx Cinachyrella apion x Geodia gibberosa x Geodia papyracea XX Xx Chondrilla nucula xX Xx Suberites zeteki x Lissodendoryx isodictyalis XX Biemna spp. NOK Clathria schoenus p Xx Clathria venosa x Mycale laevis p % Mycale aff. magniraphidifera OX Mycale microsigmatosa X xX Tedania ignis XX 24 Xx Scopalina ruetzleri Xx Ki Amorphinopsis sp.1 XX Halichondria magniconulosa? | xx i Halichondria melanodocia x Halichondria poa? XX Haliclona caerulea x Xx Halclona curacaoensis x Haliclona implexiformis XX x Halclona manglaris nox Haliclona mucifibrosa XX Haliclona pseudomolitba XX Haliclona tubifera BOX X Amphimedon erina p Calyx podatypa Xx Hyrtios proteus Xx Spongia tubulifera Xx Dysidea etheria Pp x Xx Relative abundance at Twin Cays is expressed as: xx = more common in Twin Cays than in the Pelicans; x = as common, or more common, in the Pelican Cays; p = present in Twin Cays, but not common. Relative abundance at other Caribbean locations, if indicated by the cited author, is expressed as: XX = very common, X = common. Key to locations: Bel =Twin Cays, Belize (this study); Bah = Bimini, Bahamas (Riitzler, 1969); Cub = Cuba (Alcolado, 1980); Jam = Port Royal, Jamaica (Hechtel, 1965); PaG=Galeta, Panama (Wulff, personal observation); PaS = San Blas, Panama (Wulff, personal observation); VeM = Parque National Morrocoy, Venezuela (Diaz et al., 1985); VeB = Bahia de Buche, Venezuela (Sutherland, 1985). XX XX XX ~*~ xX a eK MK XX XX Xx xX 255, 256 Table 2. The most abundant 30 sponge species on mangrove roots in the Pelican Cays, Belize. Most abundant mangrove sponges in Pelican Cays Cinachyrella apion Chondrilla nucula Anthosigmella varians Placospongia intermedia Spirastrella mollis Terpios manglaris Tethya actini Monanchora arbuscula Desmapsamma anchorata Artemisina melana Clathria schoenus Clathria venosa Mycale laevis Mycale microsigmatosa Totrochota birotulat Scopalina ruetzleri Haliclona curacaoensis Amphimedon compressa Amphimedon erina Xestospongia carbonaria Xestospongia proxima Hyrtios proteus Ircinia felix Spongia obscura Spongia tubulifera Dysidea etheria Chelonaplysilla aff. erecta Halisarca caerulea Aplysina fulva Verongula rigida Also found in Twin Cays XX x Pp xO K'UO XX “x 'S Also found on very shallow reefs in Panama XX XX Note: Relative abundance of these species in Twin Cays, Belize, is indicated by: xx = very common in Twin Cays as well as in the Pelicans; x = more common in the Pelican Cays than in Twin Cays; p = present, but not common in Twin Cays. Indication is provided of which species are also common (x) or very common (xx) on very shallow reefs (less than 3 m) in San Blas, Panama (Wulff, 1984; Clifton, et al., 1996; personal observation). Doe Table 3. Relative abundances of facultative and specialist sponge feeders and their predators in three locations in each of the Twin Cays and Pelican Cays, Belize Twin Cays Pelican Cays Sponge feeders and predators HDA CUD SPOW I CATIIMAN (EIS Potential Sponge Predators Angelfish, Pomacanthus spp. XX XX XX Trunkfish, Acanthostracion spp. x x Filefish, Aluterus scriptus Xx Spadefish, Chaetodipterus faber Parrotfish, Sparisoma spp. x x xe x Scarus isertii XX x Starfish, Oreaster reticulatus x x XX XX OK Potential Predators of Sponge Predators Barracuda, Sphyaena barracuda Xx IX XX Note: xx = very abundant, x = present and readily seen without extensive searching, but not especially abundant. Key to locations within Twin Cays and Pelican Cays: HID = Hidden Creek, CUD = Cuda Cut, SPO = Sponge Haven, CAT = Cat Cay, MAN = Manatee, FIS = Fishermans Cay. Table 4. Number of times individuals of the starfish Oreaster reticulatus were observed to be feeding on 11 species of sponges and various other foods in Twin Cays and Pelican Cays, Belize. Sponges and other starfish food Twin Cays Pelican Cays Sponges Chonarilla nucula 2 Lissodendorxy isodictyalis 1 Clathria echinata Mycale laevis lotrochota birotulata Tedania ignis Amorphinopsis sp. 2 Xestospongia carbonaria Oceanapia sp. Ircinia spp. Aplysina fulva Other invertebrates Compound ascidian l Zoanthid l Agaricia tenuifolia 12 Plant material Heavily epiphytized Halimeda sp. ] 4 Filamentous mangrove root epiphytes l Filamentous algae on peat ] Microalgae in sediments 9 Note: Three of the observations on Mycale laevis were in the form of unambiguous scars from recent (i.e., within the previous couple of days) feeding, rather than direct feeding observations, and 7 of the observations on Agaricia tenuifolia were in the form of feeding scars. NO ee — oO KS Ne WN WD 258 Mycale laevis had fallen among a dense community of sponges that are not among those preferred by these large starfishes (primarily Chondrilla nucula), and this sponge, evidently protected by these inedible species completely surrounding it, was one of the few fallen individuals of edible sponge species that were not consumed by the starfishes. Sponge species represented among individuals that had fallen off the mangrove roots or that were growing on substrates under the roots, but had not been consumed by starfishes, included Spongia tubulifera, Amphimedon erina, Chondrilla nucula, Placospongia intermedia, and Ircinia strobilina. In the Pelicans, I observed several starfishes with broken spines indicating that they had recently strayed into areas in which parrotfishes forage (Wulff, 1995b). Three bites were observed as they were being taken, all by sponge-feeding grey angelfishes. Chondrilla nucula suffered two of these bites and the other was on Aiolochroia crassa. DISCUSSION What Is a “Typical Caribbean Mangrove Sponge Community”? The more common species at Twin Cays appear to coincide with the more or less typical mangrove-associated sponge fauna (Table 1) described in the literature (Alcolado, 1990; Hechtel, 1965; Sutherland, 1980; Riitzler, 1969; Diaz et al., 1985) or personally observed (Galeta and San Blas, Panama). There are some exceptions (de Weerdt et al., 1991), however: Haliclona caerulea is not present, but the Twin Cays fauna include several other Chalinid sponges that are absent from other mangrove faunas. Also lacking are Halichondria melanodocia, Suberites zeteki, and Geodia gibberosa, which have been reported to live on mangrove roots in at least three other places in the Caribbean (Table 1). Another six species (Clathria schoenus, Mycale laevis, Amphimedon erina, Dysidea etheria, Mycale microsigmatosa, and Scopalina ruetzleri) reported on mangrove roots in at least three other Caribbean locations are present at Twin Cays, but are more common in the Pelicans. An additional nine species that are common at Twin Cays but have not been reported from other locations outside Belieze, may in large part reflect the intensity of the collecting at Twin Cays, as well as the fact that some team members recently revised important groups of mangrove sponges. In any case, the Twin Cays mangrove-root sponge community easily falls within the range of variation in membership reported among other Caribbean mangrove root sponge communities (see references in Table 1). In contrast, all but 4 of the 30 most common species in the Pelicans are closer in composition to the species inhabiting very shallow reef areas in San Blas, Panama, which are characteristic of reef habitats (Table 2). However, 16 of these 30 species appear to be absent from the nearby mangroves in Twin Cays (i.e., 80% of the species in common with a shallow reef, but fewer than 50% of the species in common with a nearby set of mangrove cays). Although several other excellent species lists of Caribbean coral reef sponge assemblages have been published (e.g., Alcolado, 1979, 1990; Alvarez et al., 1990; Schmahl, 1990), I compared the Pelicans only with Panama, in large part because of the flux in Caribbean sponge systematics in the past 10 years and thus the possible inconsistency in the names applied to the same sponges species. Because the data from Panama are my own (Wulff, 1994: Clifton et al., 1996), this at least 259 provides a consistent interpretation of the systematics and a consistent intensity of surveying. The Panama reef sponge assemblages are also in much shallower water than reef sponges reported in other studies, and this depth range makes the habitat more comparable to the mangrove habitat in at least that variable. In the San Blas Islands in Panama, coral reef sponges live in dense communities in as little as 2 m (e.g., Wulff, 1994, 1995a). On the Pelican Cays mangrove roots, these sponges can also occur in shallower water, even within the top 1 m. Exposure during extreme low tides demonstrates how shallow these sponges are and indicates one constraint on sponge distribution on mangrove roots (Riitzler, 1995). Another important constraint on sponge distribution in mangroves is sediment. Alcolado (1990) mentioned complete lack of sponges on mangrove roots in areas of high sediment, for example, near-rivers and within estuaries, and this is also the pattern in San Blas, Panama (personal observation). Wulff (1995a) suggested that the distribution of sponges in shallow water is prevented by rough water movement, whereas in Panama hurricanes virtually never occur, and during the single recorded hurricane, sponge populations were more protected on the leeward sides of reefs. The occurrence of dense communities of reef sponges in very shallow water in the Pelican Cays corroborates that shallow water per se does not restrict reef sponge distribution. Demonstrated Predator-Imposed Constraints on Habitat Distribution of Caribbean Benthic Organisms Experiments and observations of herbivorous grazers on reefs have repeatedly confirmed that herbivores influence the abundance and also the species composition of algal communities (e.g., reviews in Lubchenko and Gaines, 1981; Hay, 1997). Halos around patch reefs in seagrass meadows are one obvious indication of the importance of shelter to herbivorous fishes and sea urchins (e.g., Ogden et al., 1973) in the context of safe access to feeding areas. Many studies have also demonstrated that plant community composition can be influenced by nearby habitats in which the microtopography is complex enough to allow herbivorous fishes to find shelter from their predators, although topographic relief is not the sole important factor in some cases (e.g., Lewis, 1985). That predators exert similarly profound effects on distribution and abundance of sponge species is a relatively new idea. Although sponges have been reputed to be inedible, and in fact demonstrated to be largely inedible for potential predators that have continuous access to them, sponge predators can effectively prevent edible sponges from sharing their habitat (Wulff, 1994, 1995b, 1997a, 1997b). Sponges that do not have adequate inherent defenses against a particular predator are absent from habitats frequented by those predators. As previous studies of predation on sponges have demonstrated, some sponge species that are normally entirely hidden in crevices in the reef (e.g., Halichondria cf. lutea [possibly = Amophinopsis sp. 1 in this study] and Geodia cf. gibberosa) are readily consumed by herbivorous parrotfishes when experimentally removed from their refuges (Wulff, 1988, 1997a). These cryptic species were able to grow beyond the confines of their cryptic spaces when protected by small cages. Two sponge species that live partially hidden (Adocia sp. and Mycale laevis ) are also readily consumed by these herbivores when their surfaces are removed, indicating that defenses in these species are concentrated in their surfaces (Wulff, 1997a). If defenses are expensive to produce, concentration of defenses in surface tissue would be particularly adaptive for species that are also somewhat protected by their partly hidden habitat. Sponges living on mangrove roots are not normally challenged by sponge- 260 feeding fishes because these fishes are consistently associated with coral reefs, possibly because of the availability of shelter from their predators, which is lacking among the mangrove roots. When a Halichondria species, Geodia gibberosa, Tedania ignis, and Chondrosia collectrix were removed from mangroves and placed in a reef-fish habitat, they were also consumed, with angelfishes preferring C. collectrix and parrotfishes preferring G. gibberosa (Dunlap and Pawlik, 1996). In this study, large chunks of sponge were presented in the open on racks, however, and it is not clear that the results can be applied to the natural situation of many dozens of species growing together in dense multispecies clusters. Transplantation experiments will be required to determine if fish predators are indeed excluding the more typical mangrove sponge fauna from Pelican Cay mangroves because fishes can consume an entire edible sponge within minutes, and even removal of a single bite can leave barely a trace because the sponges heal so quickly. The influence of starfishes on sponge distribution is more readily observed. The large starfish Oreaster reticulatus extrudes its stomach and digests the sponge tissue, leaving behind distinctive feeding scars providing evidence of its meal for some days afterward (depending on the amount of spongin in the skeleton). The effect of starfish predation on sponge distribution was obvious in the Pelicans, where sponges demonstrated to be edible to O. reticulatus (Wulff, 1995b) were abundant on mangrove roots only 0.5 m or less above hard substrate, on which a very different sponge community, composed solely of species demonstrated to be rejected by the starfishes, was thriving; that is, the distribution pattern appeared to be enforced by starfish predation, because the only factor that differed between the substrates was accessibility to starfish grazing. Are Predators the Primary Influence on Sponge Community Structure in the Twin Cays and Pelican Cays? If sponge predators are restricting sponges without inherent defenses from living in the Pelicans, why is the fauna there not merely a depauperate version of the Twin Cays sponge fauna? That is, why is it not the same typical mangrove root fauna, but lacking the undefended sponge species? The far more diverse sponge fauna of the Pelicans is reminiscent of other, less complex, situations in which predators have been demonstrated to increase diversity by feeding on organisms that otherwise are capable of outcompeting many of the species in the system (e.g., Paine, 1966). Although little is known of sponge energetics, recent work by Uriz et al. (1995) suggests that allocations to secondary chemistry used in predator defense could decrease growth rates or reproductive rates. It is possible that this is the key, and decreased growth and reproduction rates may be the trade-off for increased predator resistance. Reef sponges, which would be extinguished without inherent defenses against predators, may be outcompeted on mangrove roots because they divert resources to predator defenses and thus have lower growth rates. The observation of many sponges brooding abundant larvae in populations in which recruitment by sexually generated larvae is too low to be observed (Wulff, 1991), except after a hurricane scoured the substrate (Wulff, 1995a), suggests that the limiting step in successful recruitment is not availability of larvae, but rather availability of suitable substrate. The usual mangrove sponge fauna may even be selected for increased allocation to reproduction in response to the finite nature of individual mangrove roots and the need to colonize fresh roots because expansion to adjacent space is not possible by vegetative means of propagation, as it is on the 261 reef. In his study of sponge community dynamics on mangrove roots in Venezuela, Sutherland (1980) concluded that, at least for the Venezuelan mangrove sponge fauna he studied, recruitment was a relatively rare event, followed by long periods of relative stasis. Recruitment of reef sponges has also been demonstrated to be relatively rare, except after a storm cleared the substrate (Wulff, 1995a) or in cryptic spaces near adult sponges (Zea, 1993). Growth has been demonstrated to be relatively slow and highly variable for most reef sponges (e.g., Hoppe, 1988; Wulff, 1990, 1991) and relatively rapid for at least a couple of mangrove sponge species (Ellison et al., 1996). When apparent competition between sponges has been investigated over long periods or experimentally, the intimate associations have actually been shown to be of mutual benefit to participating species in some cases (e.g., Sara, 1970; Riitzler, 1970; Wulff, 1996). Examples of mutual benefit have all come from sponge species that consistently coexist. It is possible that the distinct separation of Caribbean sponges into a typical mangrove root fauna and a typical reef fauna results in part from competitive exclusion of reef sponges from mangroves by faster-growing mangrove sponge species, except in unusual circumstances, as when the habitat context favors residence of spongivores. ACKNOWLEDGMENTS Special thanks go to Klaus Riitzler for carrying out his vision of providing an opportunity for Caribbean sponge specialists to collaborate in the field and for gathering funds and organizing logistics to bring us all together; and also to fellow workshop participants Kate Smith, Rob van Soest, Sven Zea, Cristina Diaz, Belinda Alvarez, and Klaus Riitzler for continuing inspiration. Contribution No. 587, Caribbean Coral Reef Ecosystems Program, National Museum of Natural History, Smithsonian Institution. REFERENCES Alcolado, P. M. 1979. Ecological structure of the sponge fauna in a reef profile of Cuba. In Biologie des Spongiaires, edited by C. Levi and N. Boury-Esnault, 297-302. Collog Int CNRS, Parissve 29 1990. General features of Cuban sponge communities. In New Perspectives in Sponge Biology, edited by K. Riitzler, 351-357. Washington, DC: Smithsonian Institution Press. Alvarez, B., M. C. Diaz, and R. A. Laughlin 1990. The sponge fauna on a fringing reef in Venezuela, II: Composition, distribution and abundance. In New Perspectives in Sponge Biology, edited by K. Riitzler, 358-366. Washington, DC: Smithsonian Institution Press. Clifton, K. E., K. Kim, and J. L. Wulff 1996. A Field Guide to the Reefs of Caribbean Panama with an Emphasis on Western San Blas. Proceedings of the 8th International Coral Reef Symposium, Panama. Diaz, H., M. Bevilacqua, and D. Bone 1985. Esponjas en Manglares del Parque National Morrocoy. Fondo Editorial, Acta Cientifica Venezolana. 262 Dunlap, M., and J. Pawlick 1996. Video-monitored predation by Caribbean reef fishes on an array of mangrove and reef sponges. Marine Biology 126:117-123. Ellison, A. M., E. J. Farnsworth, and R. R. Tilley 1996. Facultative mutualism between red mangroves and root—fouling sponges in Belizean mangroves. Ecology 77:2431—2444. Hay, M. E. 1997. The ecology and evolution of seaweed-herbivore interactions on coral reefs. Proceedings of the 8th International Coral Reef Symposium, Panama, 1:23-32. Hechtel, G. J. 1965. A Systematic Study of the Demospongiae of Port Royal, Jamaica. Peabody Museum of Natural History, Yale University, Bulletin 20. Hoppe, W. F. 1988. Growth, regeneration, and predation in three species of large coral reef sponges. Marine Ecology Progress Series 50:117—125. Lewis, S. M. 1985. Herbivore abundance and grazing intensity on a Caribbean coral reef. Journal of Experimental Marine Biology and Ecology 87:215—228. Lubchenco, J., and S. D. Gaines 1981. A unified approach to marine plant-herbivore interactions. I. Populations and communities. Annual Review of Ecology and Systematics 12:405—437. Ogden, J. C., R. A. Brown, and N. Salesky 1973. Grazing by the echinoid Diadema antillarum Philippi: Formation of halos around West Indian patch reefs. Science 182:715—717. Paine, R. 1. 1966. Food web complexity and species diversity. American Naturalist 100:65—76. Randall, J. E. and W. D. Hartman 1968. Sponge-feeding fishes of the West Indies. Marine Biology 1:216—225. Riitzler, K. 1969. The mangrove community, aspects of its structure, faunistics and ecology. In Lagunas Costeras, 515-536. UNAM-UNESCO, Mexico, DF. 1970. Spatial competition among Porifera: Solution by epizoism. Oecologia 5:85—95. 1995. Low-tide exposure of sponges in a Caribbean mangrove community: P.Z.N.I. Marine Ecology 16:165-179. Riitzler, K., and I. C. Feller 1996. Caribbean mangrove swamps. Scientific American 274:94—99. Sara, M. 1970. Competition and cooperation in sponge populations. Symposium of the Zoological Society of London 25:273-284. Schmahl, G. P. 1990. Community structure and ecology of sponges associated with four Southern Florida coral reefs. In New Perspectives in Sponge Biology, edited by K. Riitzler, 376-383. Washington, DC: Smithsonian Institution Press 263 Sutherland, J. 1980. Dynamics of the epibenthic community on roots of the mangrove Rhizophora mangle, at Bahia de Buche, Venezuela. Marine Biology 58:75-84. Uriz, M. J., X. Turon, M. A. Becerro, J. Galera, and J. Lozano 1995. Patterns of resource allocation to somatic, defensive, and reproductive functions in the Mediterranean encrusting sponge Crambe crambe (Demospongiae, Poeciloscerida). Marine Ecology Progress Series 124:159-170. Weerdt, W. H. de, K. Riitzler, and K. P. Smith 1991. The Chalinidae (Porifera) of Twin Cays, Belize, and adjacent waters. Proceedings of the Biological Society of Washington 104:189-205. ° Wilkinson, C. R., and E. Evans 1989. Sponge distribution and abundance across Davies Reef, Great Barrier Reef, relative to location, depth, and water movement. Coral Reefs 8:1—7. Wulff, J. L. 1988. Fish predation on cryptic sponges of Caribbean coral reefs. American Zoologist 28:166. 1990. Patterns and processes of size change in Caribbean demosponges of branching morphology. In New Perspectives in Sponge Biology, edited by K. Riitzler, 351-357. Washington, DC: Smithsonian Institution Press. 1991. Asexual fragmentation, genotype success, and population dynamics of erect branching sponges. Journal of Experimental Marine Biology and Ecology 149:227-247. 1994. Sponge-feeding by Caribbean angelfishes, trunkfishes, and filefishes. In Sponges in Time and Space: Biology, Chemistry, Paleontology, edited by R. W. M. van Soest, T. M. G. van Kempen, and J-C Braekman, 265-271. Rotterdam: A. A. Balkema. 1995a. Effects of a hurricane on survival and orientation of large, erect coral reef sponges. Coral Reefs 14:55-61. 1995b. Sponge-feeding by the Caribbean starfish Oreaster reticulatus. Marine Biology 123:313-325. 1996. Mutualisms among species of coral reef sponges. Ecology 78:146—159. 1997a. Parrotfish predation on cryptic sponges of Caribbean coral reefs. Marine Biology 129:41-52. 1997b. Causes and consequences of differences in sponge diversity and abundance between the Caribbean and eastern Pacific at Panama. Proceedings of the 8th International Coral Reef Symposium, Panama, 2:1377—1382. Zea, S. 1993. Recruitment of demosponges (Porifera, Demospongiae) in rocky and coral reef habitats of Santa Marta, Colombian Caribbean: P.Z.N.I. Marine Ecology 14:1—21. 1994. Patterns of coral and sponge abundance in degraded versus still healthy reefs at Santa Marta, Colombian Caribbean. In Sponges in Time and Space, edited by R. W. M. van Soest, T. M. G. van Kempen, and J-C Brakeman, 257—264. Rotterdam: Balkema. ieiih, © m7 ATOLL RESEARCH BULLETIN NO. 478 GNATHOSTOMULIDA IN THE PELICAN CAYS, BELIZE BY WOLFGANG STERRER ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 7 Dangriga Belize City Coco Plum Cay Be Guatemala Honduras Narco War Gay Le Twin Cays Blue Ground EUS » v 40: South Water Cay Sittee Point vai “| Carrie Bow Cay iTt.- 4 Or “* Curlew Bank qi ae ous oe ae °9 i VOUNasS Oe ae ai +0) Patch Reefs x ae & Sand Bores eehic . ue South Cut Riversdale “2 Pelican Cays 2° l G Manatee Cay . iQ Cat Cay ase oe 72. Sie! Gish oe Jonathan Point fa a Be ates, Patch Reefs “>, Oe * & Sand Bores * Cn False Point 16°35'N 5 km | 88°15'W Figure 1. Index map showing sample sites. GNATHOSTOMULIDA IN THE PELICAN CAYS, BELIZE BY WOLFGANG STERRER' ABSTRACT Gnathostomulida, a small phylum of microscopic, sand-dwelling marine worms, appear to be particularly well represented in the coarse yet detritus-rich sediments that typically occur where coral reefs meet seagrasses or mangroves. Of 25 species encountered in 35 sediment samples collected in southern Belize between 1974 and 1997, 14 species were extracted from only two samples from the Pelican Cays. INTRODUCTION Gnathostomulida are unsegmented, acoelomate, small (1-3 mm long) worms that live in the interstices of marine sand. First described as aberrant Turbellaria (Ax, 1956), they were subsequently promoted to the rank of phylum (Riedl, 1969; Sterrer, 1972) on the basis of a unique combination of anatomical features, particularly a monociliary epidermis (each epidermal cell carries only a single cilium), and a bilaterally symmetric pharynx equipped with complex cuticular mouth parts. Found exclusively, sometimes in large numbers, in shallow, detritus-rich marine sand, Gnathostomulida are presumed to feed by grazing on the bacterial and fungal microflora that coats sand grains. In addition to very low oxygen requirements, they may have mechanisms for sulfide detoxification. Only 91 species, in 25 genera, are currently known worldwide (Sterrer, 1998), many with cosmopolitan distribution. Gnathostomulida may be the most primitive living Bilateria (Ax, 1986; Sterrer et al., 1985), yet their phylogenetic affinities remain enigmatic. Between 1974 and 1997, I visited the Carrie-Bow Cay field station in southern Belize (Riitzler and Macintyre, 1982) six times and collected 88 sediment samples, which yielded a total of 25 species of Gnathostomulida (including 7 species and 2 genera new to science), the largest number from any area in the world (Sterrer, 1998). Most samples came from the immediate vicinity of Carrie-Bow Cay (see Fig. 1), either from shallow sand between patch reefs and Thalassia or from deep sand troughs in the fore-reef area. The nearby mangrove island of Twin Cays was sampled repeatedly. Sand from the base of the Southern Sandbores, curious cone- shaped islets within the lagoon, produced the type specimens of Clausognathia suicauda Sterrer, 1992. 'Bermuda Natural History Museum, Flatts FLBX, Bermuda; e-mail wsterrer@sargasso.bbsr.edu. METHODS The method of collecting and specimen extraction (for details see Sterrer, 1998) allows some crude observations on species richness (Sterrer, 1971). Using snorkeling or scuba, the upper 5 cm of sediment are scooped into a bucket by hand until the latter is full; a primary sample thus consists of about 10-15 liters of sand, with a little overlying seawater. In the lab, this primary sample is periodically subsampled by scooping the superficial layer of sand (about 500 ml) into a flask and shaking it in an isotonic magnesium sulfate solution. The floating meiofauna is then poured through a 63 «vm sieve and allowed to recover before it is sorted to species and analyzed under the phase-contrast microscope. Extraction ends when the sample ceases to produce gnathostomulids, usually after 7—12 days. RESULTS All samples were collected with the specific objective of finding Gnathostomulida, which means they came from sheltered sandy (not muddy) bottoms with a high content of marine (not terrigenic) detritus, as is typically found in the vicinity of seagrasses, mangroves, and between coral reefs. In their preferred environment, Gnathostomulida are often represented by up to a dozen species per sample, whereas marginal environments (i.e., sand that is either too clean or too clogged with terrigenous detritus) may contain only one species, often of the eurytopic genus Gnathostomula (unpublished observations). Representative morphological types of these worms are shown in Figure 2. Of a total of 88 samples taken between 1974 and 1997 (Table 1), 35 samples (40%) were positive, each containing at least one, and at most 10, species. Over the period May 5—16, 1994, I collected 24 samples ( 6 at Coco Plum Cay, 5 at Twin Cays, 2 at Southern Sand Bores, | at Man- o’ War Cay, 7 off Carrie-Bow Cay, and 3 at Pelican Cays). Of these, 14 samples (58%) yielded Gnathostomulida. The following three samples were collected in the Pelican Cays: #94.9: coll. May 7, 1994, off Cat Cay; small sand hole in reef flat off red mangrove, 2 m #94.14: coll. May 12, 1994, off Cat Cay; heterogeneous Halimeda sand from two sand troughs in reef ridge, 2m #94.15: coll. May 12, 1994, at S end of Manatee Cay; medium to coarse sand in sparse, short Thalassia, 1.5 m. Two of these samples (94.9 and 94.15) were positive (67%), yielding 8 and 10 species, respectively, or a total of 14 species. This high species richness per positive Pelican Cays sample compares favorably with the remaining 12 positive samples of the 1994 collecting season, which together produced only 7 species (0.58 spp. per sample). Of 25 species recorded from Belize, three (Prerognathia alcicornis, P. crocodilus, and P. ctenifera) were found exclusively in the Pelican Cays samples. The remarkable species richness of Gnathostomulida in the Pelican Cays may be due to the composition of the sediment—coarse coral sand admixed with flocculent detritus—itself the result of the close proximity of coral reef, seagrasses, and mangrove ("mangreef"). — epidermal cells Filospermoidea (Haplognathia rosea) Scleroperalia (Gnathostomula peregrina) Conophoralia (Austrognathia microconulifera) 0} basal plate Figure 2. Representative Gnathostomulida of the order Filospermoidea, and the suborders Scleroperalia and Conophoralia. Size range of these interstitial worms is 1-3 mm long. 270 Table 1. Comparison of Gnathostomulida sampled at The Pelican Cays and other locations in Belize between 1974 and 1997. Pelican Cays, 1994 Other Belize __—_ Belize Taxa and Parameters 94.9 94.15 total 1994 _ total total Order Filospermoidea Family Haplognathiidae Haplognathia asymmetrica Sterrer, 1991 0 3 3 0 3 6 belizensis Sterrer, 1998 0 0 0 0 2 2 lunulifera (Sterrer, 1969) 0 0 0 0 1 1 rosea (Sterrer, 1969) 0 0 0 3 2 12) ruberrima (Sterrer, 1966) 1 1 2 0 13 15 Family Pterognathiidae Cosmognathia aquila Sterrer, 1998 0 2. 2 6 18 20 arcus Sterrer, 1991 0 0 0 0 1 l manubrium Sterrer, 1991 0 l ] 0 1 2 Pterognathia alcicornis Sterrer, 1998 l 0 l 0 0 1 crocodilus Sterrer, 1991 1] 0 1 0 0 1 ctenifera Sterrer, 1969 0 1 7 0 0 i swedmarki Sterrer, 1966 0 0 0 0 1 1 ugera Sterrer, 1991] 9 0 9 0 6 15 Order Bursovaginoidea Suborder Scleroperalia Family Clausognathiidae Clausognathia suicauda Sterrer, 1992 | 0 l 0 10 11 Labidognathia longicollis Riedl, 1970 0 ] l 0 7 8 Tenuignathia rikerae Sterrer, 1976 0 0 0 ] 17 N7/ Family Paucidentulidae Paucidentula anonyma Sterrer, 1998 0 l | 0 6 7 Family Onychognathiidae Onychognathia rhombocephala Sterrer, 1998 0 0 0 1 15 15 Family Gnathostomulidae Gnathostomula axi Kirsteuer, 1964 0 0 0 0 11 1] peregrina Kirsteuer, 1969 l l 2 0 10 12 Suborder Conophoralia | Family Austrognathiidae Austrognathia christianae Farris, 1977 0 2 2 0 19 21 microconulifera Farris, 1977 1 1 2 2 10 12 Austrognatharia medusifera Sterrer, 1998 | 0 0 0 3 8 8 sterreri (Kirsteuer, 1969) 0) 0 0 0 8 8 strunki Farris, 1973 0 0 0 8 8 8 Total positive samples 1 1 2 12 33 35 Total specimens IS 20 S55 24 187 222 Total species 7 10 14 7 747} 25 Species per sample | 7.00 10.00 7.00 0.58 0.67 0.71 271 ACKNOWLEDGMENTS I am indebted to Klaus Riitzler and Mike Carpenter for providing such congenial field facilities on the doorstep to near-pristine marine ecosystems. Fieldwork for this project was supported by the U.S. National Museum of Natural History's Caribbean Coral Reef Ecosystems Program (CCRE Contribution No. 588). REFERENCES Ax, P. 1956. Die Gnathostomulida, eine ratselhafte Wurmgruppe aus dem Meeressand. Abhandlungen der Akademie der Wissenschaften und Literatur Mainz, mathematisch- naturwissenschafiliche Klasse, 8:1—32. 1986. The position of the Gnathostomulida and Platyhelminthes in the phylogenetic system of the Bilateria. In The Origins and Relationships of Lower Invertebrates, edited by S. Conway Morris, J. D. George, R. Gibson, and H. M. Platt, 168-180. Oxford: Clarendon Press. Riedl, R. J. 1969. Gnathostomulida from America. Science 163:445—-442. Riitzler, K., and I. G. Macintyre (eds.) 1982. The Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize, 1. Structure and Communities. Washington, DC: Smithsonian Institution Press, 539 p. Sterrer, W. 1971. Gnathostomulida: Problems and procedures. Smithsonian Contributions to Zoology 76:9-15. 1972. Systematics and evolution within the Gnathostomulida. Systematic Zoology 21(2):151-173. 1992. Clausognathiidae, a new family of Gnathostomulida from Belize. Proceedings of the Biological Society of Washington 105(1):136—142. 1998. Gnathostomulida from the (sub)tropical northwestern Atlantic. Studies on the Natural History of the Caribbean Region 74:1-178. Sterrer, W., M. Mainitz, and R. M. Rieger 1985. Gnathostomulida: Enigmatic as ever. In The Origins and relationships of lower invertebrates, edited by S. Conway Morris, J. D. George, R. Gibson, and H. M. Platt, 183-199. Oxford: Clarendon Press. ier ll iS ferret a ha Civ, > ? 7 Ther inin le ATOLL RESEARCH BULLETIN NO. 479 ECHINODERMS OF THE RHOMBOIDAL CAYS, BELIZE: BIODIVERSITY, DISTRIBUTION, AND ECOLOGY BY GORDON HENDLER AND DAVID L. PAWSON ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 = Caribbean Mexico = f /itaninouse ie Rio Roti Aviéennia Cay 4 natee Cay *.*. Zapotilla Cays oS F Honduras _——) 50 km sist PERE. :[plittle Cat Cay +>. |88'w eo Pelican Cays *- 5 : ae * 26m z ();Quamino Cay So ‘ Wee Cay: : Seo Cay ™. ot a ~ { Reef ERE UY Northeast Cay p , 17°N (Ge a ion e Fd on . s Dangriga ae Glover's Reef Bird cos 7a z : 18m LI Carrie Bow Cay : eaeee Cay : a ?Gladden Spit a8 oe ae i 2 Co Cat eave 4 Yt Fisherman's:Cay L a 3 ay \— F 0 s Bs oe 25m : (/chennel Cay ’ 27m 25m i Omi : : Se East Cut “ea Lagoon Cays 4): & i en Or 26m "90m ","s ‘ d hon vey! i Slasher : = oe ‘oe 20m #2 Crdcken, Sand “0. wae Bore Craw! Cay ‘ rg Boe S 31m 33m Victoria Channel Inner Channel Bakers Rendevous 2 Bec eS ee Taropum Cay Round Cay Carbonate shoals (<2m depth) Mangrove Cays Lay 2km |s8°w Figure 1. Map of Pelican Cays and the surroundin echinoderms. g rhomboidal cays that were surveyed for ECHINODERMS OF THE RHOMBOIDAL CAYS, BELIZE: BIODIVERSITY, DISTRIBUTION, AND ECOLOGY BY GORDON HENDLER' and DAVID L. PAWSON” ABSTRACT Fifty-two species of echinoderms were found in a preliminary survey of 13 sites in the Pelican Cays and the nearby rhomboidal cays in the southern region of the Belize Barrier Reef lagoon. Most are a subset of the 86 species known from the barrier reef and offshore atolls. More species of echinoderms are associated with coral and rubble on the shelf and slope around the cays than in the bays and ponds. Some echinoderms may be excluded from the cays by the low diversity of corals and consequent lack of habitat complexity, the lack of solid substrate and the reduced water flow in protected embayments, and physical stresses, including extreme temperatures and salinities. Ten species found at the cays have not previously been reported from Belizean waters. Among them are Ocnus suspectus and Thyone pseudofusus, the first dendrochirote sea cucumbers reported from Belize. It is suggested that the cays offer suitable substrates, calm waters, and possibly a refuge from predation for some species that are cryptic on, or completely excluded from, reef habitats. Ophioderma cinereum and Ophioderma appressum of unusually large size and dark coloration adopt exposed positions on mangrove peat banks, whereas in reef habitats they are cryptic. Echinometra viridis, which shelters beneath rubble in turbulent reef habitats, takes exposed positions at the cays. Its putative role in shaping the composition of coral communities such as the cays is discussed. Ophiophragmus pulcher and Amphipholis cf. januarii, the first long-armed burrowing amphiurid brittle stars reported from Belize, and Ophiopsila riisei, typically found in rocky reef substrates, are associated with soft, peat bank substrates. At the cays, the first instance of Oreaster reticulatus feeding on a living coral was noted. Its prey, Agaricia tenuifolia, is a dominant species on the steep slopes of the cays. Since Oreaster recruits to seagrass beds on the cays, it may be a potential threat to the long-term stability of local coral communities. Synaptula hydriformis is reported to be the only echinoderm present in a pond on Elbow Cay, where its success may depend on its capabilities for colonization, its lack of planktonic larvae, and its physiological tolerance of environmental extremes. The pond population of S. hydriformis might be a relict from the soft-bottom and coral community that formerly occupied the site or may be of recent origin. The species is a viviparous, self-fertilizing hermaphrodite and potentially a proficient colonist. ‘Natural History Museum of Los Angeles County, Los Angeles, CA, 90007, USA. *National Museum of Natural History, Smithsonian Institution, Washington, DC 20560, USA to ~~] lon) INTRODUCTION Two notable features of the Pelican Cays, in the southern part of the Belize Barrier Reef lagoon, are the remarkably high species diversity of certain taxa and the abundance of some typically rare and cryptic organisms (Goodbody, 1995; Littler and Littler, 1997). Their unusual ecology is attributed to: the occurrence of coral, mangrove, and seagrass communities in intimate proximity; oligotrophic lagoon habitats with an ample circulation of oceanic seawater; and minimal sediment resuspension. The sensitive habitats probably have been preserved thanks to negligible human visitation to these islands. Studies of the Pelican Cays habitats are warranted because they have a unique and poorly understood ecology, they are highly susceptible to degradation through natural and anthropogenic disturbance, and a better understanding of the cays is needed to promote their conservation (Goodbody, 1995). Thus, in 1997 Klaus Riitzler, Director of the Smithsonian Institution’s Caribbean Coral Reef Ecosystems (CCRE) program, invited the authors to conduct a survey of the Pelican Cays echinoderms in order to bolster efforts to preserve their habitats. The present report describes our findings for 1997 and our prior observations in the Pelican group and surrounding cays in 1986 and 1990, which were made in connection with a more extensive survey of the Belizean echinoderm fauna that is in progress. We summarize the published records for the distribution of Belizean echinoderms and contrast them with our results for the Pelican Cays. In addition, we discuss the species composition and noteworthy characteristics of the echinoderms occurring at the Pelican group. Information is also presented concerning the surrounding islands, including Elbow, Lagoon, Quamino, Bakers Rendezvous, and Tarpum cays because all, like the Pelican group, are parts of “atoll-like” rhomboidal shoals (Fig. 1). Throughout this report, the latter cays and the Pelican group are referred to collectively as the rhomboidal cays. METHODS As indicated in Table 1, eight sites in the Pelican group and five neighboring cays were surveyed for echinoderms in 1986, 1990, and 1997. One or two localities were examined at each island. Before 1997, the cays were reached by inflatable boats and whalers that were too small to carry passengers and scuba gear for long distances, and so observations were made using snorkel. Even in calm seas, the trip from Carrie Bow Cay to Elbow Cay took 45 minutes and almost 3 hours to the Lagoon Cays, limiting the time available for fieldwork. In 1997, several localities in the Pelican Cays and vicinity were surveyed with a larger, faster fiberglass boat, permitting scuba observations on several island slopes. Nevertheless, the diversity of habitats examined and the intensity of sampling were limited by time constraints and by efforts to minimize destructive sampling. Aerial photographs of the cays were taken on March 15, 1989, in the late morning, from an altitude of approximately 500 m. Most of the species discussed herein are conspicuous epifauna or are associated with large pieces of rubble or small corals that could be easily overturned. Soft bottom and peat bank substrates were sampled sporadically when signs of infauna were noted. In some cases, samples of algae and sponge that were carefully examined and pieces of rubble that were accidentally collected yielded species of echinoderms that would otherwise have been overlooked. However, DISCUSSION Composition of the Echinoderm Fauna on the Rhomboidal Cays and Barrier Reef Belize has a rich shallow-water echinoderm fauna with more than 90 species (Table2; representative species, Fig. 7). The present study suggests that the greatest diversity, comprising approximately 86 species, occurs on coral reef habitats of the Barrier Reef and offshore atolls. Fewer species are associated with the cays of the Barrier Reef lagoon. Most are a subset of the reef-associated fauna, although some have not been found on the reef. A distinctive suite of species is restricted to soft-sediment benthic habitats near the mainland. Only a few mainland species have been reported, which may be attributed to a lack of sampling along the coast and to the influence of terrigenous sediments, river runoff, and other environmental factors. However, several species listed herein for the Barrier Reef environs—including Paraster doederleini, Moira atropos, and Brissopsis elongata—were found exclusively in a mud field within the lagoon (see Kier, 1975) and might better be categorized with the mainland group once their distribution is charted. Kier (1975) found Paraster cf. Paraster floridiensis in the mud field, and tests of the species were found in shallow water at Bird and Little Cat cays in this study, suggesting it may be more eurytopic than the other mud flat spatangoids. In the present study, 7 echinoderm species that had not been found on barrier reef and atoll environments were collected from the rhomboidal cays. Kier (1975) reported several sea urchins, listed in Table 2 as belonging to the barrier reef and environs (Lytechinus variegatus, Arbacia punctulata, and Clypeaster rosaceus), as occurring exclusively in the Thalassia beds east of Twin Cays, a lagoonal mangrove cay. The occurrence of the same 3 species at the rhomboidal cays, and the 7 species found exclusively at the cays, suggests that there is a distinctive Belizean mangrove cay fauna composed of at least 10 species. However, it is not readily apparent why the species are absent from the barrier reef. The 7 species that were found only at the rhomboidal cays have all been reported from reef and seagrass habitats elsewhere in the Caribbean; only Echinaster echinophorus has previously been reported from mangrove habitats (Hendler et al., 1995). Ocnus suspectus, Thyone pseudofusus, and Pseudothyone belli, which were found in the rhomboidal cays, represent the first Belizean records of holothuroids in the order Dendrochirotida. Pawson (1976) noted that the absence of dendrochirote sea cucumbers in collections from Carrie Bow Cay was inexplicable, given the importance of the group at other Caribbean localities. Thus the discovery of Belizean dendrochrotes in this study is not surprising, but the reason for their apparent restriction to the cays in the barrier reef lagoon 1s an enigma. It may be significant that 7. pseudofusus is unusual among dendrochirotes in that it 1s a facultative deposit-feeder. It is not known how O. suspectus and P. belli feed. As noted above, 10 species present at Pelican Cays localities were not found at the other 5 thomboidal cays surveyed, and 10 species found at the other rhomboidal cays were absent from the Pelican group. The discrepancies are probably an artifact of sampling, since many of the species are small and easily overlooked, most of them were represented at only one sampling station, and some occurred in substrates that were not systematically sampled. With 44 species of echinoderms, the Pelican group scarcely has greater species richness than the surrounding rhomboidal cays, where 43 species were found. There are similarities and striking differences between the ophiuroid species that were 278 Table 1. Localities and dates for surveys of echinoderms at the Pelican group and surrounding cays. Collecting mode, approximate time devoted to the survey, and station numbers are provided for each collection. Location 7 Apr 86 10 Jun 90 15 Jun 90 Elbow Cay Northeast Cay Bird Cays Fisherman’s Cay Manatee Cay Manatee Cay Shoal Cat Cay East Cut Little Cat Cay Lagoon Cays Quamino Cay Tarpum Cay Bakers Rendezvous Cay CBC86-24 CBC90-2 (bay) snorkel (pond) CBC86-25 CBC86-26 snorkel 1 hr CBC86-27 (outer bay) snorkel 4 hr snorkel CBC90-12 (bay) snorkel CBC90-13 (pond) snorkel CBC90-10 snorkel CBC90-11 snorkel Survey Dates 5 Sep 97 CBC97-4 snorkel | hr CBC97-5 scuba 12 m 1 hr CBC97-3 scuba 18 m | hr 8 Sep 97 CBC97-15 scuba 15 m 1 hr CBC97-12 Snorkel | hr CBC 97-14 snorkel CBC97-13 scuba 18 m 1 hr 10 Sep 97 11 Sep 97 CBC97-19 (bay) snorkel CBC97-24 (Frenchy’s Ponds, inner cove) CBC97-25 (outer cove) Snorkel 1 hr CBC97-26 (Great Bay) snorkel CBC97-21 CBC97-22 (slope) scuba 18 m | hr CBC97-23 (inner bay) snorkel 279 substrates that serve as microhabitats for small cryptic species were not systematically examined. When habitats were resampled, novel species were found, and others seen previously at the locality were overlooked, indicating that the sampling methodology was not altogether exhaustive. RESULTS The Distribution of Belizean Echinoderms Echinoderm species from Belize reported in reliable sources and those found in the present study are listed in Table 2. The sole records for the mainland of Belize are for Belize Harbor and Stann Creek (= Dangriga) (John and Clark, 1954; Kier, 1975). Information for Turneffe Islands, Lighthouse Reef, and Glover’s Reef atolls is based on Devaney (1974) and Boone’s (1928) report of Meoma ventricosa from Glover’s Reef. Records for the barrier reef are derived from surveys at Carrie Bow Cay and environs for echinoids (Kier, 1975), holothuroids (Pawson, 1976), crinoids (Macurda, 1982), ophiuroids (Hotchkiss, 1982; Hendler, 1984, 1988, 1995; Hendler and Miller, 1984; Hendler and Littman, 1986; Hendler and Turner, 1987; Hendler and Peck, 1988; Hendler et al., 1995), and asteroids (Miller, 1984). Additional species of Belizean echinoderms were noted by Hendler et al. (1995), but only their records for specific locales (e.g., Carrie Bow Cay) are considered here. A total of 88 shallow-water species were previously recorded (Table 2). Three species are reported only from the mainland: Leodia sexiesperforata, Luidia clathrata, and Amphipholis gracillima. The latter two have not been found elsewhere in Belize. Thirty-one species are reported for the offshore atolls, and Astropecten duplicatus has been found only there. The richest locale, and the most intensively sampled, is Carrie Bow Cay and environs, where 83 species are known. Twenty-eight species were found there exclusively. They include species limited to the deep fore reef (Ophiobyrsa serpens, Ophionereis vittata, Ophioderma ensiferum, Sigsbeia conifera), species associated with reef-dwelling hosts (Ophiactis quinqueradia, Ophiothrix lineata), and a suite of species typical of exposed Caribbean reef habitats (Crinoidea, Copidaster lymani, Poraniella echinulata, Ophioderma anitae, O. guttatum, O. phoenium, Echinometra lucunter), as well as three species of Leptosynapta that are endemic to Carrie Bow Cay. The survey of the 13 rhomboidal cays sites yielded 53 species. Forty-four species were found at the Pelican Cays localities, and 43 species on the nearby reefs. Eight species occurring on the cays had not been found at the barrier reef or offshore atolls: Echinaster echinophorus, Ophiophragmus pulcher, Amphipholis cf. januarii, Asterina folium, Ocnus suspectus, Thyone pseudofusus, Pseudothyone belli, and Holothuria floridana. The latter 6 species are new records for Belizean waters. Ten species found at the Pelican Cays were not found at the 5 rhomboidal cays nearby, and 10 species found at the latter cays were not collected in the Pelican group. General Characteristics of the Cays and of the Echinoderms at Each Site This section provides a description of each cay (see Fig. 1) and of the echinoderms based on field observations. The echinoderm species occurring at each cay are listed in Table 2. 280 Table 2. Distribution of the echinoderm species found in Belizean waters. Data for the mainland coast (M), offshore atolls including Turneffe Islands, Lighthouse, and Glover’s Reef atolls (A), and the Belize Barrier Reef in the vicinity of Carrie Bow Cay (BR) are based on literature cited in Results. Data for the cays surveyed in the present study are listed from north to south: Elbow Cay (EL), Northeast Cay (NE), westernmost of the Bird Cays (B), Fisherman’s Cay (FM), Manatee Cay (MN), Manatee Cay shoal (MS), Cat Cay (CA), East Cut (EC), Little Cat Cay (LC), northernmost of the Lagoon Cays (LG), Quamino Cay (Q), Tarpum Cay (TP), and Bakers Rendezvous Cay (BR). X indicates species for which voucher specimens were collected, # indicates non-vouchered records from field notes, and ? denotes field identifications deemed questionable. Echinoderm Species M A BR EL NE B FM MN MS. CA EC L@ULGROMMERED Class Crinoidea Order Comatulida Family Comasteridae Davidaster rubiginosus* WD # Davidaster discoideus xX Family Colobometridae Analcidometra armata x Family Antedonidae Ctenantedon kinziei x Class Asteroidea Order Paxillosida Family Luidiidae Luidia clathrata xX Family Astropectinidae Astropecten duplicatus xX Order Valvatida Family Asterinidae Asterina folium Xx Family Ophidiasteridae Copidaster lymani X Linckia guildingti xX xX Family Asteropseidae Poraniella echinulata Xx Family Oreasteridae Oreaster reticulatus* X # # X # X # X # Se OX Order Spinulosida Family Echinasteridae Echinaster (Othilia) echinophorus X xX eK Class Ophiuroidea Order Phrynophiurida Family Gorgonocephalidae Astrophyton muricatum* X DK OK Family Ophiomyxidae Ophioblenna antillensis xX X X Xx Ophiobyrsa serpens Xx Ophiomyxa flaccida SG Oh DS 20 Oh OK X KO KO Xa xX Order Ophiurida Family Ophiuridae Ophiolepis gemma X xX Ophiolepis impressa Se DE DG DE OS Xx KS XS EX xX Ophiolepis paucispina xX XxX Ophiolepis pawsoni xX Family Ophiocomidae Ophiocoma echinata eee SD 28 X xX X Ophiocoma paucigranulata Dike dak Ophiocoma pumila We DR DK X xX X # Ophiocoma wendtii 2 OX DK DDS 2K OK x Ophiocomella ophiactoides Xx 281 Table 2.-continued Echinoderm Species MEAG BREECSNER BE IEMEIMN IMS ACA EC VLG GO) SIP BA Ophiopsila riisei xX X x X a OK OK PDK Xx Family Ophionereididae Ophionereis olivacea Ophionereis reticulata xX Ophionereis squamulosa Ophionereis vittata Family Ophiodermatidae Ophioderma anitae Ophioderma appressum* Xx Ophioderma brevicaudum xX Ophioderma brevispinum Ophioderma cinereum xX Ophioderma ensiferum Ophioderma guttatum Ophioderma phoenium x Ophioderma rubicundum x Ophioderma squamosissimum Ophiurochaeta littoralis X Family Hemieuryalidae Sigsbeia conifera Family Ophiactidae Ophiactis algicola Ophiactis quinqueradia Ophiactis savignyi xX Family Amphiuridae Amphiodia pulchella Amphipholis gracillima X Amphipholis ct. januarii xX Amphipholis squamata Amphiura fibulata Amphiura stimpsonii xX Ophiophragmus pulcher X xX Ophiostigma isocanthum # Ophiostigma siva Family Ophiotrichidae Ophiothrix angulata Ophiothrix lineata Ophiothrix orstedii Ophiothrix suensontit Class Echinoidea Order Cidaridae Family Cidaridae Eucidaris tribuloides WS Dak # # Order Diadematoida Family Diadematidae Diadema antillarum MS 1 OK # Order Arbaciidae Family Arbaciidae Arbacia punctulata xX XxX xX Xx Order Temnopleuroida Family Toxopneustidae Lytechinus variegatus KX Lytechinus williams X Xx Tripneustes ventricosus X X x KKK PK mK OK KK OO OK ~< xm mK ~*~ A Se x -K mK ~< ~~ x KK mK KKK MK th ~ ~*~ * ~*~ ~ Order Echinoida Family Echinometridae Echinometra lucunter Xx Echinometra viridis* KT OKI XL # xX # # xX x wm BS DS 282 Table 2.-continued Echinoderm Species M__A_ BR EL NEBL EM MNEIMS CAS EGR CRE GROMER EEA Order Clypeasteroida Family Clypeasteridae Clypeaster rosaceus Clypeaster subdepressus Family Mellitidae Leodia sexiesperforata xX X Order Cassiduloida Cassidulus cariboearum X Order Spatangoida Family Schizasteridae Agassizia excentrica Moira atropos X Paraster doederleini Paraster cf. P. floridiensis Family Brissidae Brissopsis elongata X Brissus unicolor Meoma ventricosa xX X Plagiobrissus grandis Xx Class Holothuroidea Order Dendrochirotida Family Cucumariidae Ocnus suspectus X Family Sclerodactylidae Euthyonidiella destichada xX Pseudothyone belli x Family Phyllophoridae Thyone pseudofusus xX Order Aspidochirotida Family Stichopodidae Isostichopus badionotus X X X # xX Xe ex Isostichopus macroparentheses X Family Holothuriidae Actinopyga agassizi X ? ? ?. X Holothuria (Cystipus) cubana X xX X Holothuria (Halodeima) floridana xX x K Holothuria (Halodeima) mexicana* Si # xX X x Holothuria (Thymiosycia) arenicola X Holothuria (Thymiosycia) impatiens xX X 2 xX X xX Holothuria (Thymiosycia) thomasi xX X X Xx Order Apodida Family Synaptidae Euapta lappa X Leptosynapta imswe Leptosynapta nannoplax Leptosynapta roseogradia Synaptula hydriformis Family Chirodotidae Chirodota rotifera *See Figure 7 x ~ a) x xa KKM ~ mK mK mH KM YO ~*~ ~ <>< > x“ 283 Elbow Cay. This elongate island is at the northern tip of an extensive rhomboidal shoal that lies west of the Pelican group (Fig. 2). Its large stands of mangroves partly encircle several bays and completely enclose a small pond (Figs. 2, 4). Features of the island—its deeply indented shoreline, bays with protective sills, and enclosed pond—are similar to those of the Pelican Cays. In aerial views the outline of the island resembles the reticulate shape of the neighboring rhomboidal shoals. The appearance of Elbow and Pelican cays suggest that mangroves, once established on a rhomboidal shoal, may form bays by overgrowing shallow portions of the reef, and may form ponds by obstructing the mouths of bays. In time, the bays could give rise to ponds, and ponds could give rise to bays or fill with mangrove swamp, depending on the advance and retreat of the mangroves and corals that colonize the shoals. As noted below, corals and burrowing and boring bivalves are buried in the unconsolidated sediment in the Elbow Cay pond. They are apparently the remnants of a now defunct soft-bottom coral community, which is consistent with the view that the site was formerly an open embayment (Fig. 5). The sites examined were the large, west-facing and small, east-facing bays and the pond near the northern end of the island. The western bay is approximately 15 m deep, with a sill across its mouth reaching within less than | m of the surface. The seaward slope of the sill is covered with rubble, and colonized by calcareous algae and gorgonians. Within the bay the slope is covered with coarse sediment and a little Thalassia. Shallow, hard-bottom areas within the large bay have small coral heads scattered over the surface, including Siderastrea, clumps of branching Porites and Millepora, along with gorgonians, rubble, Thalassia, and sediment with many Halimeda flakes. Areas of exposed mangrove peat have a cover of seagrass, algae (including Padina), sponge, and some patches of Halimeda flakes. In places where the peat banks are steep and undercut, there are free-hanging mangrove roots covered with sponge, Clavelina, Caulerpa and Lobophora, and a few with large, stacked plates of Porites astreoides. Near the south edge of the large bay a sandy ridge rises to within 2.5 m of the surface. Porites rubble and sponge (Spheciospongia?) is present on the top of the ridge. Seagrass grows on the crest and on both flanks of the ridge. Spadefish, grunts, and snapper were seen in the large bay. The eastern bay is 9 m deep and is separated from the larger bay by a shallow sill. The hard bottom nearby has scattered corals and sponge, and the sand bottom had Thalassia and Syringodium. The reef slope seaward of the east bay had a Thalassia cover and a few large sponges (Spheciospongia?). The enclosed pond, at most 5 m deep, is separated from the large bay by a stand of mangrove approximately 30 m wide (Fig. 3). The bottom of the pond is covered with unconsolidated sediment about 0.6 m thick, composed largely of microscopic fecal pellets of unknown origin. Near the center of the pond there is a thinly covered framework of dead coral, the bivalves Chione cancellata and Gastrochaena hians, and peat (Fig. 5). Visibility in the pond ranges from 3 to 5 meters horizontally. The murkiness is due to suspended fine particulates, the mixing of water with different densities, and possibly the presence of tannins. Temperature and salinity measurements indicate that the exchange of water between the pond and sea and mixing within the pond are limited. After a brief rainstorm on June 10, 1990, refractometer readings indicated the surface salinity in the pond was 31 ppt and the bottom salinity was 36 ppt. On June 15, 1990, the water temperature at a depth of 0.3 m inside the pond was 32°C, whereas in the bay it was 30°C. An open channel to the reef was not seen, but cooler and warmer parcels of water were felt by Hendler while snorkeling in the pond. See Figure 2. Elbow Cay: oblique aerial view showing the small eastern bay, larger western bay. and the enclosed pond at the northern end of the island. Reef surrounding the island forms a shallow sill at the mouths of the bays and joins a rhomboidal shoal which is seen to the south of the cay. Figure 3. Elbow Cay: dense stand of Rhizophora mangle surrounding the enclosed pond. 285 ‘puod pasojsua ay} JO 19]U99 dy) WO] YjJaIUDI auolyD SATRAIG SY} PUL STRIOD Jo sJUoWIsRIy ‘ABD MOQTA *¢ DNB ‘OUITJIDIM JY} 1B JOP Saaz] aaoisueW pol JO sjoo1 dosd ay} ul paysurjus oed]e usasd snoyusweyy JO seul YOIY], “M2IAIOAO ‘puod pasojous :AeD MOI “p aNSIy hs ae Sg 286 At the edge of the pond there are floating mats of filamentous algae (Fig. 4), and mangrove roots thickly covered by Caulerpa verticillata and other algae, dangling a meter into the water. Other conspicuous biota associated with the roots include extremely numerous sabellids and actinarians, various sponges and gastropods, with a bullid snail in abundance, oysters, a green colonial ascidian, tanaids, and shrimps. The soupy sediment is covered with patches of white microbial mat near the shore, and elsewhere with a thin layer of greenish material (cyanobacteria and/or diatoms?) supporting a rich meiofauna. Barracuda, snappers, silvery schooling fish, and mosquitofish were seen in the pond. The peat banks in the western bay have a distinctive echinoderm fauna, including Ophiopsila riisei, a long-armed brittle star that typically nestles in coral interstices at reef habitats. Hendler et al. (1995) stated that it lives in mangrove peat banks after observing a dense population of the species in the large bay at Elbow Cay in 1986, and again in 1997. Individuals do not extend their arms during the day and are presumably active nocturnally. Those excavated from peat occupy hollow cavities such as those found in decomposing twigs, with their arms enfolding the disk. The occurrence of two species of small nestling amphiurid brittle stars is also notable. Amphipholis squamata occurs in a wide range of microhabitats, and Amphiodia pulchella is usually found on the reef in soft sediment or in association with algae and corals (Hendler and Littman, 1986; Hendler et al., 1995). Ophioderma cinereum, Ophiolepis impressa, and Ophiomyxa flaccida were found under clumps of Millepora and sponge, and Ophioderma brevispinum in peat and mangrove leaves and under rubble amid seagrass. An orange sponge, Mycale laevis, yielded 249 specimens of Ophiactis savignyi. They were nearly all small and fissiparous; only one among the few large individuals had five arms. In 1990, Echinometra viridis was noted to be much more common than Lytechinus variegatus. In 1997, at the same locality, small L. variegatus were particularly abundant under rubble; E. viridis was again common and was sometimes found under the same pieces of rubble as Arbacia punctulata. The only echinoderm species seen in the Elbow Cay pond is Synaptula hydriformis. In Bermuda the species is usually mottled brown and white, and green and white individuals have been noted at Carrie Bow Cay (Pawson, 1976, 1986). The living individuals at Elbow Cay are brown and turn red when preserved in ethanol. Northeast Cay, Pelican Cays. The southwest part of the cay has an exposed sandy coastline with Thalassia and Syringodium near the shore. Barnacles grow on mangrove roots along the shore. In the bay nearby, cover for echinoderms is provided by rubble, sponge, a small amount of Acropora cervicornis, Millepora, zoanthid mats, gorgonians, and Padina. Clumps of Agaricia can be seen along the crest and flank of the slope outside the bay, but the lower part of the slope is sandy. Mounds of conch shells and clumps of Halimeda occur in the water near a fish camp. One Lytechinus variegatus was seen in the bay, with its peristomial region and viscera missing, but with the spines on the test still moving. It is not known if the urchin was preyed upon by a boxfish seen nearby. Bird Cays, Pelican Cays. Large brain corals were seen at the collecting site on the westernmost part of the Bird Cays. Echinoderms were collected from rubble, Mil/epora, and coral occurring in the shallows near the island, which were dominated by Thalassia. 287 Fisherman’s Cay and Ponds E and F, Pelican Cays). Some details can be added to a previous description of the habitats at Fisherman’s Cay (Littler and Littler, 1997). Ctenophores were observed carried by the current into the larger (outer) Pond E_ but the water was relatively still in the smaller (inner) Pond F. Evidently, it is only at the mouths of the two ponds that water currents sweep the bottom clean and expose firm sediment. Pen shells are embedded in sediment near the entrance to the outer pond. Elsewhere the sediment is soupy, with white microbial patches in places. Pendent mangrove roots support huge masses of Caulerpa verticillata and C. racemosa, Halimeda opuntia, ascidians, various sponges, and an abundance of small sabellid worms. A dense bed of Thalassia is present in the outer cove. The sessile fauna is strikingly different on the hard shelf immediately outside Frenchy’s Ponds E and F. There, large gorgonians and sponges, and large plates of coral (Montastrea or Porites?), grow on Rhizophora roots. The steep shelf slope is densely populated with sponge, gorgonians, and living Agaricia. The northern side of the cay, at the mouth of Great Bay, is the site of a broad shoaling Thalassia flat with many clumps of Millepora and Porites, some overgrown by zoanthids. The wall of the seaward slope has a cover of coral rubble and sponge. Oreaster reticulatus was the only echinoderm seen in Pond E. In the field, a superficial examination of algae in the inner cove was unproductive, but a one-gallon sample of Halimeda opuntia from the outer cove studied in the laboratory yielded two specimens of Ophiothrix angulata. The undercut peat banks might have harbored echinoderms, but a cursory search for them was unproductive. In contrast to the pond, the cay’s shelf and slope support a profusion of Ophiothrix suensonii, as well as Isostichopus badionotus, Holothuria mexicana, and Lytechinus variegatus. Echinometra viridis is abundant at Great Bay. Echinaster echinophorus occurs in the open on sandy algae-covered sediment, and Ophiopsila riisei is common in the clumps of algae growing on the top of the slope at the mouth of the embayment. Manatee Cay, Pond C, and Manatee Cay Shoal, Pelican Cays. In Pond C, the surface water is very warm and a reddish color. Mangrove roots on the northern side of the pond are heavily overgrown with sponge, algae, and colonial tunicates. In places the peat banks are 2 m high and strongly undercut. Sponges cover the ceiling of the undercuts; patches of decomposing mangrove leaves cover areas of the soft-sediment floor. French angel fish and barracuda were seen in the pond. A dive was made on both sides of a shoal at the south end of the island. Acropora cervicornis was found growing sparsely on the summit of the ridge. A dense cover of Agaricia fouled by sponges and fine sediment, together with some large coral shelves (Montastrea annularis?), is present on both flanks. One section of the shoal is free of living coral and thickly covered with rubble. In Pond C, burrowing amphiurid brittle star arms can be seen on the surface of the mud. Ophiophragmus pulcher and Amphipholis cf. januarii were collected from soft sediment beneath the overhanging peat banks. They were the only echinoderms seen in Pond C. In the seagrass bed surrounding the island, Ophioderma appressum and O. cinereum were collected from beneath coral rubble and large Holothuria mexicana and Oreaster reticulatus were observed in the open. The latter was also found on the reef slope, and an individual was seen with its arms clasping a colony of Agaricia tenuifolia. Beneath the extruded stomach of the sea star, tissue of the coral was dissolving and white skeleton was exposed. Echinometra viridis was found to be common on the shoal, on coral, and on rubble-dominated portions of the slope. Actinopyga agassizi and 288 Holothuria mexicana were collected from the rubble-covered area, and a small individual of Asterina folium was collected on a piece of rubble gathered with the sea cucumbers. Cat Cay and East Cut, Pelican Cays. The wall on the northeast slope of the cay and the southern side of the opening at East Cut were observed while using scuba. The latter site is a break near the middle of a shoal that extends southward for approximately three kilometers from Cat Cay. Off Cat Cay, at depths of 4 to 6 m, there is a narrow shelf with rubble cover and small corals, large brain corals, and Rhipocephalus brushes. The steep wall of the slope consists of silty sediment colonized by colonies of Agaricia tenuifolia, plates of large agariciids, and shingled brain corals, along with some Halimeda, delicate sponges, gorgonians, and antipatharian wires. The slope at East Cut is similar, with rubble overgrown by Lobophora, Agaricia tenuifolia, agariciids, and scattered gorgonians and antipatharian wires being the predominant cover. A brief boat reconnaissance at Channel Cay, due east, indicated that coral cover is sparser there than at East Cut. The echinoderm fauna on the shallow shelf is fairly diverse; exposed Oreaster reticulatus is present, along with Ophioderma spp., Ophioblenna antillensis, and Ophionereis reticulata associated with corals and rubble. Echinometra viridis is very abundant on the slope, and Ophiothrix suensonii and Ophiocoma wendtii are fairly common. At a depth of 9 m, a juvenile Astrophyton muricatum was found entwined in a gorgonian. The only ophiuroids seen at depths >9 m were large Ophioderma cinereum and Ophioderma rubicundum. E. viridis was not recorded at East Cut, but Arbacia punctulata was collected there. Little Cat Cay, Pelican Cays. Thalassia grows close to the mangrove roots in shallow water at the southeast corner of the cay. A reef flat with coral, rubble, and sponges occurs where the bottom slopes gradually to a depth of approximately 1 m. A long shoal extending southward from the island has a steeply sloping western flank reaching more than 15 m in depth. Hard- bottom areas in the bight on the southwest side of the island are composed of densely packed Porites rubble. Mangroves roots are embedded in a thin layer of peat and sand. A ridge crest occurs several meters from the mangrove shoreline. It is occupied primarily by clumps of Agaricia and Millepora, some overgrown with algae, sponge, or zoanthid, and by some thickets of Acropora cervicornis. Sandy slumps occur between the coral patches. The sandy slope of the ridge is strewn with rubble, scattered sponges, and gorgonians. It drops at a 45° angle, to a depth of more than 15 m, onto a gently sloping sand bottom. Acropora cervicornis on the ridge crest is densely covered with climbing Echinometra viridis at one site, and at another the branches are overgrown with a greenish-brown sponge. Clumps of Agaricia were found to harbor Ophiomyxa flaccida, Ophioderma appressum and O. rubicundum, and Ophioblenna antillensis; Ophiothrix angulata and Ophionereis reticulata were found in Halimeda. Many large blue and yellow individuals of Ophiothrix suensonii were conspicuous on the slope of the ridge. Ophiopsila riisei was found at Little Cat Cay, but in less abundance than at Elbow Cay. A small individual of Astrophyton muricatum at the site had an associated shrimp, Periclimenes perryae. Three large individuals of Diadema antillarum were noted on a peat bank; this was the only sighting of the species in rhomboidal cays. Lagoon Cays. The northerly cay (Fig. 6) was first surveyed in 1986, and the small outer cove of the bay was examined at that time. The reef flat was occupied by gorgonians with dead 289 branch tips that may have been killed by exposure, and by coral heads of moderate size. A bed of Thalassia with short blades grew on a densely packed coral rubble bottom that was covered with a thin layer of sand. The bed extended to the shore, where mangrove roots were fixed in the sand and there was a very shallow peat bank. In 1997, seagrass was noted on the southeast coast of the island, outside the bay. Figure 6. Lagoon Cays: oblique aerial view showing the larger northern cay, smaller southern cay, and surrounding shoals. The northern cay has an enclosed pond, which was not explored, and a bay with a large, deep cove and a small shallow cove. In 1997, we examined the deep inner pond of the bay. The mangrove roots hanging from high, steep peat banks were thickly covered with oysters, and abundant Ascidia nigra and other solitary tunicates with associated sabellid worms. At that time, 5% of the 4. nigra were turning white and disintegrating. Beneath the bank, terebellid tentacles extended from the soft sediment and in other areas there were accumulations of Halimeda flakes. Seaward of the bay there was a silty slope on the flank of the cay, with little coral cover. On the northeast side of the cay, near the crest of the reef, the shelf at 6 m depth is dominated by seagrass, with gorgonians, clumps of Porites, moderate-sized coral heads, and a few large brain corals. The reef slope is covered with soft sediment and scattered living Agaricia 290 and agariciid plates, gorgonians, and antipatharian wires. At 12 m depth, the cover is provided by algae; much of the Agaricia there is dead. At a depth of 18 m, the bottom is composed largely of silty soft sediment, except for occasional coral clumps and Halimeda. At Lagoon Cays, the presence of Ophioderma cinereum and Ophioderma appressum in exposed positions on the tops of high peat banks is noteworthy, since both species are invariably cryptic wherever they occur in reef habitats (Hendler et al., 1995). Most of the individuals were large. Among those collected were an O. cinereum (disk diameter, 35.7 mm; arm length, 163 mm) and an O. appressum (disk diameter, 25.3 mm; arm length, 125 mm) approaching the maximum size known for the two species (Hendler et al., 1995). Orange-red individuals of Echinaster echinophorus were found in the open on the walls of the peat bank. The arms of burrowing amphiurids were not observed in or near the bank, but individuals of Ophionereis reticulata extended their arms from the sediment below the banks and anchored their disks in the branches of buried Porites rubble. Ophiothrix angulata was perched on top of epiphytic Ascidia nigra and was found along with Ophiopsila riisei in the shallow mangrove peat bank near the outer part of the bay. Several species of holothuroids, Ophionereis olivacea, Ophiocoma echinata, and Arbacia punctulata, which were collected in 1986 but not in 1997, were found amid rubble near the outer part of the bay. The density and abundance of echinoderms on the reef slope was greatest near the crest and decreased downslope. At 18 m only a few individuals of Ophioderma and Ophiocoma were found associated with coral rubble. However, Amphiura stimpsonii, Ophiothrix angulata, Ophiactis algicola, and juvenile Ophiopsila riisei, Ophiostigma cf. siva, and Echinometra viridis were found in a sample of Halimeda. A full-grown specimen of O. siva was found on a piece of rubble collected along with holothuroids. Most species occurring in the inner bay—including Echinometra viridis, Ophioderma cinereum, O. appressum, O. reticulata, Ophiopsila riisei, Ophiolepis impressa, and Ophiothrix angulata—are a subset of those collected on the reef slope at the Lagoon Cays. However, Lytechinus variegatus, Echinaster echinophorus, and Isostichopus badionotus are restricted to the bay. Ophiomyxa flaccida, Ophioderma rubicundum, Ophiocoma wendtii, Ophiostigma siva, Ophiothrix suensonii, O. oerstedii, Ophiactis algicola, Actinopyga agassizii, and Holothuria impatiens only occur on the reef slope. Quamino Cay. Quamino Cay occupies the northeast end of a narrow rhomboidal shoal that borders the inner channel of the Belize barrier reef lagoon and lies to the west of Little Cat Cay. During a brief snorkeling reconnaissance, a Thalassia bed was seen abutting the mangroves on the eastern edge of the island. A shallow shelf extending from the northeast corner of the cay has considerable sponge and algae cover, Siderastrea and heads of brain coral, clumps of Agaricia, and sparse Acropora cervicornis. Whenever a piece of rubble was turned, a dense cloud of fine sediment spread through the water. The reef slope also has a thick cover of silty sediment and scattered Agaricia and gorgonians. Most of the echinoderms collected at Quamino Cay live in dead clumps of Agaricia and in large pieces of rubble, which are overgrown with algae. The most common species on the shallow shelf are Ophioderma cinereum, Ophionereis reticulata, and Ophiocoma pumila. The latter two typically nestle in rubble-covered sediment at back-reef habitats such as Carrie Bow Cay. 291 Tarpum Cay. The cay is at the northern end of a group of rhomboidal shoals to the southeast of the Pelican Cays. Observations made from the boat indicated there was sparsely distributed Agaricia on the silty slope around the island. The survey was made while snorkeling in the island’s south-facing bight. There, peat banks along the mangrove-lined shoreline are shallow and very hard. Exposed roots bear meager algae and sponge cover and, in pools between the roots, there are accumulations of decomposing leaves and white microbial mat. The fine white sand beyond the mangroves covers mangrove roots and fallen trunks, and has a sparse growth of Thalassia and Caulerpa. Further from shore, where the depth is 1 to 3 m, the bottom is composed of fine white silt with numerous mounds and burrows, and individuals of Cassiopea are numerous. In places at Tarpum Cay, the water washing out of mangroves is an intense brownish color, presumably from tannins in solution. The site is relatively unproductive, in part because there is hardly any rubble to shelter echinoderms, and the sand bottom and peat are too hard-packed for most burrowing species. An effort to locate burrowing brittle stars in the peat banks was not successful. However, purely epifaunal organisms were found, including Holothuria mexicana and Oreaster reticulatus. Bakers Rendezvous Cay. The cay is south of Quamino Cay and southwest of Tarpum Cay and the Pelican group. It is situated on a very long narrow shoal that separates Victoria Channel from the broad Inner Channel of the barrier reef lagoon. The dive site was located on the steeply sloping face of the island to the east of the gap between the two main stands of mangroves on the cay. As at Quamino Cay, the water was murky and the bottom very silty. At a depth of 6 m, there is a shelf covered by a Thalassia bed; at greater depths on the slope Agaricia tenuifolia appears, along with some large agariciid plates, gorgonians, antipatharian wires, and a few small brain corais. Echinometra viridis was very common at this site. A single individual of Lytechinus williamsi was found, the only one seen in the rhomboidal cays. The individuals of Ophioderma cinereum (max. 23.4 mm) and Ophioderma rubicundum (max. 14.5 mm in disk diameter) that were collected were of typical size. Characteristics of the Mangrove Cay Fauna Table 3 compares the dominant ophiuroid species from the rhomboidal cays and the barrier reef. It shows that the cays fauna are more similar to that of the shallow rubble habitats on the barrier reef than to that of the living substrates in shallow water on the barrier reef and the fore-reef slope. The comparison is inexact, as it is based on the number of sites, of a possible maximum of 13, at which the rhomboidal species occur, versus an actual percentage of the population represented by each species at three Barrier Reef habitats. Among the echinoderms found at five or more localities in the rhomboidal cays were 13 ophiuroids, Echinometra viridis, Lytechinus variegatus, Oreaster reticulatus, Holothuria mexicana, H. impatiens, H. cubana, H. floridana, and Isostichopus badionotus (Table 3). The most widespread echinoderms were Ophioderma cinereum and Echinometra viridis, both of which were found at 11 localities. Ophiomyxa flaccida, Ophionereis reticulata, Ophioderma appressum and Ophiolepis impressa were almost as common; each was found at 10 of the 13 surveyed localities in the rhomboidal cays. Table 3. Frequency of occurrence of the most widespread rhomboidal cays echinoderms compared with the most numerically abundant barrier reef ophiuroids. Rhomboidal cays species are listed in the order of the number of sites (out of 13) at which each was observed. The 10 most common ophiuroids at the rhomboidal cays fall above the dashed line. The ophiuroids that were among the 10 most common at the barrier reef, and are also among the most widespread in the rhomboidal cays, are indicated by an X. Their ranking was calculated in terms of the percentage of the population each species comprised at 3 different habitats at Carrie Bow Cay (based on Hendler and Peck, 1988: 414). Rhomboidal Cays Barrier Reef (all taxa) (ophiuroids only ) Number of sites Shallow reef Shallow reef _ Fore-reef Species (descending) (in living substrata) _(in rubble) slope Ophioderma cinereum 1] Echinometra viridis Ophiomyxa flaccida 10 4 Ophionereis reticulata Xx Ophioderma appressum Xx xX Ophiolepis impressa X Oreaster reticulatus Ophioderma rubicundum 9 xX x Xx Ophiothrix suensonii Ophiopsila riisei 8 x Ophiothrix angulata x Holothuria mexicana Ophiocoma echinata 7 x Xx Lytechinus variegatus Ophiocoma wendtii 6 xX x xX Ophiothrix orstedii xX x Holothuria impatiens TIsostichopus badionotus Ophiocoma pumila 5 x x Holothuria cubana Holothuria floridana Several species of ophiuroids found at the rhomboidal cays were distinguished from their reef-dwelling conspecifics by differences in size and pigmentation. As detailed in the treatment of Lagoon Cays, individuals of Ophioderma cinereum at the rhomboidal cays are also unusually large. Reef-dwelling individuals of O. cinereum are characteristically pale gray. Those at the cays are typically brown or grayish-brown, with a tendency for small individuals to be grayish and large individuals to be dark brown. Ophiolepis impressa collected from the cays were sometimes a deeper brown color than individuals from the Barrier Reef habitat. The individuals of Ophionereis reticulata at the cays have a less distinct and contrasting netted pattern on the disk than those from many coral reefs. 293 DISCUSSION Composition of the Echinoderm Fauna on the Rhomboidal Cays and Barrier Reef Belize has a rich shallow-water echinoderm fauna with more than 90 species (Table2; representative species, Fig. 7). The present study suggests that the greatest diversity, comprising approximately 86 species, occurs on coral reef habitats of the Barrier Reef and offshore atolls. Fewer species are associated with the cays of the Barrier Reef lagoon. Most are a subset of the reef-associated fauna, although some have not been found on the reef. A distinctive suite of species is restricted to soft-sediment benthic habitats near the mainland. Only a few mainland species have been reported, which may be attributed to a lack of sampling along the coast and to the influence of terrigenous sediments, river runoff, and other environmental factors. However, several species listed herein for the Barrier Reef environs—including Paraster doederleini, Moira atropos, and Brissopsis elongata—were found exclusively in a mud field within the lagoon (see Kier, 1975) and might better be categorized with the mainland group once their distribution is charted. Kier (1975) found Paraster cf. Paraster floridiensis in the mud field, and tests of the species were found in shallow water at Bird and Little Cat cays in this study, suggesting it may be more eurytopic than the other mud flat spatangoids. In the present study, 7 echinoderm species that had not been found on barrier reef and atoll environments were collected from the rhomboidal cays. Kier (1975) reported several sea urchins, listed in Table 2 as belonging to the barrier reef and environs (Lyfechinus variegatus, Arbacia punctulata, and Clypeaster rosaceus), as occurring exclusively in the Thalassia beds east of Twin Cays, a lagoonal mangrove cay. The occurrence of the same 3 species at the rhomboidal cays, and the 7 species found exclusively at the cays, suggests that there is a distinctive Belizean mangrove cay fauna composed of at least 10 species. However, it is not readily apparent why the species are absent from the barrier reef. The 7 species that were found only at the rhomboidal cays have all been reported from reef and seagrass habitats elsewhere in the Caribbean; only Echinaster echinophorus has previously been reported from mangrove habitats (Hendler et al., 1995). Ocnus suspectus, Thyone pseudofusus, and Pseudothyone belli, which were found in the rhomboidal cays, represent the first Belizean records of holothuroids in the order Dendrochirotida. Pawson (1976) noted that the absence of dendrochirote sea cucumbers in collections from Carrie Bow Cay was inexplicable, given the importance of the group at other Caribbean localities. Thus the discovery of Belizean dendrochrotes in this study is not surprising, but the reason for their apparent restriction to the cays in the barrier reef lagoon is an enigma. It may be significant that 7. pseudofusus is unusual among dendrochirotes in that it is a facultative deposit-feeder. It is not known how O. suspectus and P. belli feed. As noted above, 10 species present at Pelican Cays localities were not found at the other 5 rhomboidal cays surveyed, and 10 species found at the other rhomboidal cays were absent from the Pelican group. The discrepancies are probably an artifact of sampling, since many of the species are small and easily overlooked, most of them were represented at only one sampling station, and some occurred in substrates that were not systematically sampled. With 44 species of echinoderms, the Pelican group scarcely has greater species richness than the surrounding rhomboidal cays, where 43 species were found. There are similarities and striking differences between the ophiuroid species that were 294 Figure 7. Common representatives of echinoderms in situ in Belize: a, Feather star, Davidaster rubiginosa (Crinoidea), top arrow, among gorgonians on slope of coral ridge near Cat Cay, 5 m; ophiuroid (lower arrow) arms protruding from lettuce coral (Agaricia) in right foreground (photo I.G. Macintyre). b, Cushion sea star, Oreaster reticulatus (Asteroidea) (photo K. Ritzler). c, Basket star, Astrophyton muricatum (Ophiuroidea) (photo C. Clark). d, Brittle star, Ophioderma appressum (Ophiuroidea) on top of sea urchin (photo G.M. Miller). e, Sea urchin, Echinometra viridis (Echinoidea) (photo G.M. Miller). f, Sea cucumber, Holothuria mexicana (Holothuroidea) (photo K. Sandved). 295 most widespread at the rhomboidal cays, and the species that are most abundant on the barrier reef at Carrie Bow Cay (Table 3). A list of the 10 most widespread ophiuroid species from the rhomboidal cays includes 8 of the most abundant species found by Hendler and Peck (1988) in 3 barrier reef habitats near Carrie Bow Cay (i.e., Ophiomyxa flaccida, Ophionereis reiculata, Ophioderma appressum, Ophiolepis impressa, Ophioderma rubicundum, Ophiopsila riisei, Ophiothrix angulata, and Ophiocoma echinata). Ophioderma rubicundum, which was the most numerous species on the fore-reef slope of the barrier reef, was relatively widespread in the rhomboidal cays. However, ophiuroid species at the barrier reef that were most numerous in shallow-water reef rubble (Ophiocoma echinata) and in shallow-water living substrates (Ophiocoma pumila) were not particularly widespread at the rhomboidal cays. Moreover, the most widespread ophiuroid at the rhomboidal cays (Ophioderma cinereum) was relatively rare on the reef (Hendler and Peck, 1988) and thus is not listed for the barrier reef in Table 3. Because so little is known about their natural history, it is difficult to explain why some species have been successful in the rhomboidal cays and others have not though they are abundant on the barrier reef (Hendler et al., 1995). It is possible that the population density and distribution of rhomboidal cays echinoderms generally hinge on coincidental, stochastic events in population recruitment and extinction. It is also possible that their success depends on crucial adaptations in reproductive mode, behavior, competitive capabilities, or physiology, and on environmental factors peculiar to the cays. The alternative explanations should be tested, since the results might provide a better understanding of the rhomboidal keys and of large-scale changes on Caribbean reefs. Presence, Absence, and Impact of Echinoderm Species in the Rhomboidal Cays E. viridis is one of the most widespread echinoderms of the rhomboidal cays. Our observation of large numbers of E. viridis climbing on the branches of Acropora cervicornis in 1986 at Little Cat Cay may have been a typical occurrence at that time, and E. viridis has persisted as a dominant herbivore throughout the rhomboidal cays (Aronson and Precht, 1997). It is replaced on reefs exposed to heavy wave action by congeneric E. /ucunter, which is capable of excavating galleries in coral rock. E. viridis typically occurs in coral and rubble on reefs, but can also be found in the open on mangrove prop roots and branching corals, and it has been suggested that its cryptic behavior may be a response to water turbulence or to predation (Hendler et al., 1995). It has been proposed that E. viridis played an important role in the drastic transformation of the rhomboidal cays reefs from a coral community that was dominated by Acropora cervicornis for at least 3,800 years, to one dominated by Agaricia tenuifolia (Aronson and Precht, 1997). Aronson and Precht indicate that the transition occurred in the 1980s because A. cervicornis succumbed to mass mortality from white-band disease. They speculate that the subsequent dominance of A. tenuifolia was abetted by E. viridis herbivory, citing Sammarco’s (1982) demonstration that FE. viridis promotes the settlement and successful growth of Agaricia spp. They contend that the Acropora-Agaricia transition occurred in Belize and not elsewhere in the Caribbean because of an exceptionally high population density of E. viridis. Aronson and Precht (1997) also suggest that its abundance could have resulted from the mass mortality of a competing species, Diadema antillarum. However, Lessios (1995) has shown that a release from competition with D. antillarum has not benefited E. viridis. Furthermore, the present-day density 296 of E. viridis in Belize is not greater than at other Caribbean localities (e.g., Lessios, 1995) and is considerably lower than the density of the echinoids monitored by Sammarco (1982) that improved the success of Agaricia spp. Whether herbivory by E. viridis controls the composition of the rhomboidal cays coral community remains to be confirmed. Oreaster reticulatus may pose a potential threat to the integrity of the Agaricia community that dominates the rhomboidal keys slopes at present. Our observation of the starfish feeding on Agaricia tenuifolia on the slope of Manatee Cay is the first evidence that O. reticulatus, which was known to have a fairly generalized diet, is a potential coralivore (Hendler et al., 1995). Agariciids, which constitute the dominant coral cover on the slopes of the rhomboidal reefs, represent a potential windfall for a coral predator. Most of the rhomboidal cays have shallow water seagrass beds, to which O. reticulatus recruits. Should there be a large, successful recruitment of the sea star, the adults would in all likelihood move downslope to feed on Agaricia, perhaps initiating another large-scale transformation of the coral community. The widespread distribution and the large body size of Ophioderma cinereum in the rhomboidal cays contrast with the characteristics of the populations at Carrie Bow Cay reef and the barrier reef. It is not clear why the species is relatively uncommon on the barrier reef, especially since it is a major component of the reef-flat ophiuroid fauna in Panama (Hendler, unpublished observation). However, the large size and the exposed position adopted by individuals on high peat banks indicate that individuals in mangrove habitats are responding to abundant resources or less predation and may be long-lived or fast-growing. It is less clear why the large individuals commonly hide in the interstices of coral and rubble on the reef slopes of several cays, but the same factors may be involved. The rarity of populations on peat banks, and the possibility that large individuals have a long life-span, are indications that experimental manipulations of those populations should be avoided. Another interesting aspect of O. cinereum is the contrast in pigmentation noted above, between individuals living on the mangrove cays and those on the reef. Presuming that the coloration is not strictly under genetic control, their integument probably incorporates pigmented material from the environment or accumulates it from their diet as they grow. Several species of burrowing brittle stars provide further examples of contrasting ecologies between reef-dwelling and rhomboidal cay populations. Large, long-armed, burrowing amphiurid brittle stars such as Ophiophragmus pulcher have not been found on the reef and seagrass habitats in Belize, although they occur in those habitats elsewhere in the Caribbean (Hendler et al., 1995). The presence of O. pulcher at Elbow Cay, and at Manatee Cay with a smaller burrowing species, Amphipholis cf. januarii, suggests that the mangrove peat banks provide a more suitable environment than the soft sediment habitats near the barrier reef. The other amphiurid brittle stars that occur on the reef are all diminutive species that do not burrow deeply. The occurrence of numerous Ophiopsila riisei burrowing in the peat banks at Elbow, Lagoon, and Little Cat cays was unexpected, as noted above, because the species is typically restricted to hard substrates on the barrier reef. Evidently, mangrove peat is a suitable substrate for some burrowing brittle stars, including O. pulcher, A. cf. januarii, and O. riisei, whether they live in hard or soft substrates elsewhere. However, the occurrence of burrowing brittle stars in the rhomboidal cays that do not occur in soft substrates at the barrier reef may be due to negligible turbulence and other environmental factors in the cays, or to the influence of biological factors such as curtailed competition and predation. Thus the factors potentially affecting the success of PESTA burrowing brittle stars in the rhomboidal cays are the same ones thought to influence E. viridis and O. cinereum. The occurrence of large numbers of Synaptula hydriformis, a small apodan sea cucumber, in the Elbow Cay pond indirectly sheds light on the absence of other species from the habitat. The pond is a small body of water with sluggish circulation and at times higher temperatures and more extreme salinities than in the barrier reef lagoon. It is likely that echinoderms rarely recruit to the pond because the surrounding mangrove swamp is a barrier to adult individuals and planktonic larvae. Quite possibly, high water temperatures and salinity in the dry season and low salinity in the rainy season are stressful to adults and lethal to larval echinoderms. Those conditions would make it difficult for echinoderms with a planktonic larval stage to maintain viable populations in the pond. The success of S. hydriformis may be due in part to its viviparous reproduction, which could protect its embryos from environmental stress. Or the population in the pond may be descended from individuals that inhabited the soft-bottom coral community, which we suggest previously occupied the same site. However, its capabilities as a self-fertilizing hermaphrodite would increase the likelihood that a single individual could successfully invade and reproduce, even in an isolated pond (Hendler et al., 1995; Frick, 1998). The situation at Elbow Cay is similar to that in Bermuda, where S. hydriformis inhabits isolated salt ponds (Pawson, 1986). The level of environmental stress in enclosed ponds must be more severe than that in the more open bays, but even in some bays of the rhomboidal cays temperature and salinity stresses probably exclude some sensitive organisms. At Elbow Cay, where only one echinoderm inhabits the pond, the nearby bay has more than 30 species. Physical stress and competition for space probably account for the absence of corals from localities such as Ponds E, F, and C. At those sites, accumulations of unconsolidated soft sediment and the absence of living corals are obstacles for the many echinoderms that require hard substrate or that are obligate symbionts of coral and other reef fauna. Thus it is not surprising that the richness of echinoderm species was markedly lower in the embayments at Manatee and Lagoon cays, in comparison with the adjacent reef slope. Even on the reef slopes at the rhomboidal cays coral diversity and spatial complexity are limited. As a result, this area has failed to attract species restricted to conditions on the deep fore- reef slope or specific to hosts living only on the barrier reef. In the absence of those species, echinoderm diversity at the cays has remained lower than on the barrier reef. The general trends in diversity that we have described are presumably reliable, although the figures for species richness at certain localities are not precise. Our findings are preliminary and limited to the more conspicuous components of the echinoderm fauna. The biodiversity of echinoderms at the rhomboidal cays is undoubtedly greater than known at present. To produce a more definitive taxonomic list of Belizean echinoderms, additional localities must be explored and their microhabitats must be examined systematically. ACKNOWLEDGMENTS Our fieldwork on echinoderms was carried out from 1979 to 1997 under the auspices of the Smithsonian Institution and with support from its IMSWE, SWAMP, and CCRE programs and Fluid Research Funds. Numerous individuals contributed to its success and deserve credit here, above all Klaus Riitzler, who paved the way for these investigations. Among the many 298 others who have our gratitude are the shipmates and dive buddies for the surveys discussed in this report, listed chronologically: Barbara Littman, Rich Aronson, Jim Lynch, James Johnson, Edd Barrows, Mike Carpenter, Cheryl Thacker, Haris Lessios, Chris Thacker, Bert Pfeiffer, and Dave Smith; also Charlie Usher who piloted an overflight of the cays. We are grateful for identifications provided by Klaus Riitzler, Steve Cairns and Lindsey Groves, and for comments on a draft of the manuscript that were offered by Florence Nishida. The Belize Ministries of Natural Resources and of Fisheries and Agriculture kindly granted permission to conduct research and to collect specimens in Belize. CCRE Contribution No. 589. REFERENCES Aronson, R. B., and W. F. Precht 1997. Stasis, biological disturbance, and community structure of a Holocene coral reef. Paleobiology 23:326-346. Boone, L. 1928. Scientific results of the first oceanographic expedition of the Pawnee 1925: Echinodermata from tropical east American seas. Bull. Bingham oceanogr. Coll. (E1225 plsai—s: Devaney, D. M. 1974. Shallow-water echinoderms from British Honduras, with a description of a new species of Ophiocoma (Ophiuroidea). Bull. Mar. Sci. 24:122—164. Bricks JE: 1998. Evidence of matrotrophy in the viviparous holothuroid echinoderm Synaptula hydriformis. Invertebr. Biol. 117:169-179. Goodbody, I. 1995. Ascidian communities in Southern Belize — a problem in diversity and conservation. Aquat. Conserv. 5:35-358. Hendler, G. 1984. The association of Ophiothrix lineata and Callyspongia vaginalis: A brittlestar- sponge cleaning symbiosis? Mar. Ecol. 5:379-401 (2) 1988. Western Atlantic Ophiolepis (Echinodermata: Ophiuroidea): A description of O. pawsoni new species, and a key to the species. Bull. Mar. Sci. 42:265-272. 1995. New species of brittle stars from the western Atlantic, Ophionereis vittata, Amphioplus sepultus, and Ophiostigma siva, and the designation of a neotype for Ophiostigma isocanthum (Say) (Echinodermata: Ophiuroidea). Contrib. Sci. (Los Angel.) No. 458:1-19. Hendler, G., and B. S. Littman 1986. The ploys of sex: relationships among the mode of reproduction, body size and habitats of coral-reef brittlestars. Coral Reefs 5:31—42. Hendler, G., and J. E. Miller 1984. Ophioderma devaneyi and Ophioderma ensiferum, new brittlestar species from the western Atlantic. Proc. Biol. Soc. Wash. 97:442461. Hendler, G., J. E. Miller, D. L. Pawson, and P. M. Kier 1995. Sea Stars, Sea Urchins and Allies: Echinoderms of Florida and the Caribbean. Washington, D.C.: Smithsonian Institution Press. 299 Hendler, G., and R. W. Peck 1988. Ophiuroids off the deep end: Fauna of the Belizean forereef slope. In Echinoderm Biology: Proceedings of the Sixth International Echinoderm Conference, edited by R. D. Burke, P. V. Mladenov, P. Lambert, and R. L. Parsley, 411-419. Rotterdam: Balkema. Hendler, G., and R. L. Turner 1987. Two new species of Ophiolepis (Echinodermata: Ophiuroidea) from the Caribbean Sea and Gulf of Mexico: With notes on ecology, reproduction and morphology. Contrib. Sci. (Los Angel.) No. 395:1-14. Hotchkiss, F. H. C. 1982. Ophiuroidea (Echinodermata) from Carrie Bow Cay, Belize. In The Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize, |. Structure and Communities, edited by K. Riitzler and I. G. Macintyre. Smithson. Contrib. Mar. Sci. No. 12:387-412. John, D. D., and A. M. Clark 1954. The “Rosaura” Expedition 1937-1938. 3. The Echinodermata. Bull. Br. Mus. Nat. Hist. (Zool.) 2:139-162, pl. 6. Kier, P. M. 1975. The Echinoids of Carrie Bow Cay, Belize. Smithson. Contrib. Zool. No. 206: iii + 1-45. Lessios, H. A. 1995. Diadema antillarum 10 years after mass mortality: still rare, despite help from a competitor. Proc. R. Soc. Lond. B 259:331-337. Littler, D. S., and M. M. Littler 1997. An illustrated marine flora of the Pelican Cays, Belize. Bull. Biol. Soc. Wash. 9:1-149. Macurda, D. B., Jr. 1982. Shallow-water Crinoidea (Echinodermata) from Carrie Bow Cay, Belize. In The Atlantic Barrier Reef Ecosystem at Carrie Bow Cay, Belize, \. Structure and Communities, edited by K. Riitzler and I. G. Macintyre. Smithson. Contrib. Mar. Sci. No. 12:413-416. Miller, J. E. 1984. Systematics of the ophidiasterid sea stars Copidaster lymani A. H. Clark, and Hacelia superba H. L. Clark (Echinodermata: Asteroidea) with a key to species of Ophidiasteridae from the Western Atlantic. Proc. Biol. Soc. Wash. 97:194—208. Pawson, D. L. 1976. Shallow-water sea cucumbers (Echinodermata: Holothuroidea) from Carrie Bow Cay, Belize. Proc. Biol. Soc. Wash. 89:369-382. 1986. Phylum Echinodermata. In Marine Fauna and Flora of Bermuda, edited by W. Sterrer, 522-541. New York: John Wiley. Sammarco, P. W. 1982. Echinoid grazing as a structuring force in coral communities: Whole reef manipulations. J. Exp. Mar. Biol. Ecol. 61:31—55. ore 6 17 > (89 : if @ee*)] itu win AGS Ret PEE Hh: azo! pee cr TN ONE > Jw , PARAS Usil's Hee dirt Mle) ihe ; mr* a ee. j 14 ma ee ers 4 hk 1 os ‘oon aa. =~ OT Ua Tis % ATOLL RESEARCH BULLETIN NO. 480 DIVERSITY AND DISTRIBUTION OF ASCIDIANS (TUNICATA) IN THE PELICAN CAYS, BELIZE BY IVAN GOODBODY ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 —————=— km 5 f \ - Northeast Cay $34 3 Mexico i o .- Se Biro Cays 4; if A Ceo Sys . a Sea ; ee D3 Belize City f og yr 7} NY Uy Dangriga y Guatemala Honduras 29m LagoonCays GS: Figure 1. Index map of the Pelican Cays showing the location of the ponds. DIVERSITY AND DISTRIBUTION OF ASCIDIANS (TUNICATA) IN THE PELICAN CAYS, BELIZE BY IVAN GOODBODY' ABSTRACT The Pelican Cays at the southern end of the Belize Barrier Reef, Central America, have a rich ascidian fauna. Seventy species in thirty genera, primarily from mangrove ponds and coral ridges, are recorded here. This number of species represents 60% of all known shallow-water ascidians in the Caribbean. The inventory includes two species normally associated with the Indo-Pacific region, Diplosoma virens and Botrylloides perspicuum. The principal habitat of each species in the Pelican Cays and the general ecology of the ascidian fauna are also discussed. INTRODUCTION The Ascidiacea (sea squirts) are soft-bodied sessile marine invertebrate animals that are common throughout the tropical Caribbean. Many species occur as solitary individuals (zooids) but others, as a result of complex asexual replicatory mechanisms, live in colonies of varying form. Some (e.g., Ecteinascidia turbinata; Plate 3e) form bushy clusters, others (e.g., Polyclinum constellatum) form cushions, while still others (e.g., species of Didemnum, Botryllus, and Symplegma) grow as flat encrusting sheets. My observations at various locations (Jamaica, Belize, Netherlands Antilles) suggest that the distribution and abundance of ascidians in the Caribbean is determined by a number of environmental factors: (1) Ascidians require a hard substratum on which the larva can settle and metamorphose and hence on which the adult zooid or colony can grow. As a result, their occurrence in the Caribbean tends to be limited to the undersurface of stones and boat hulls, and to pilings, piers, and the hanging roots of the red mangrove (Rhizophora mangle). (2) Ascidians appear to prefer shaded rather than brightly illuminated places. In Rhizophora-root communities, they tend to be abundant where the canopy overhang is dense, and on reef flats they occur mainly on the undersurface of stones except in species such as Diplosoma virens, which contain algal symbionts and grow in brightly illuminated situations. (3) In general, ascidians are absent from localities with strong wave action or surge. Thus they are uncommon on exposed reef faces and are often less abundant on the windward side of mangrove lagoons exposed to the trade winds than on the leeward side where conditions are sheltered. (4) Although dependent on suspended organic particulate matter and phytoplankton for food, ascidians appear intolerant of high concentrations of such material because it may clog their filtration system. Only a few species (e.g., Ascidia interrupta, A. sydneiensis, Polycarpa ‘Department of Life Sciences, University of the West Indies, P.O. Box 12, Kingston 7, Jamaica. 304 spongiabilis, Herdmania momus, Pyura munita, Molgula occidentalis) are regularly found living in soft substrata. When removed from the substratum, such animals are usually attached to a small shell or other hard object that the larva has apparently selected for settlement. Although the preceding factors appear to play an important role in the distribution and abundance of the ascidian fauna of the Pelican Cays, it must be remembered that these are by and large untested observations, several of which could be subjected to more rigorous scientific examination. This is particularly true in the case of the effect of light on adult ascidians. Some species have special pigmentation over the area of the neural gland and ganglion; for instance, Perophora regina has a conspicuous area of white pigment over the neural region; Corella minuta, which tends to live under stones on reef flats where it is protected from bright light, has a glassy, transparent test flecked with yellow or white pigment, but if the stone is turned over and the animal left exposed to light, pigment concentrates in a patch over the neural region (personal observations). On the other hand, Eudistoma olivaceum exhibits a small black pigment spot at the anterior end of the endostyle, suggesting that there may be some benefit in protecting this region from bright light. (For further discussion of environmental conditions affecting ascidian distribution see Kott, 1974.) The physiography of the Pelican Cays is described elsewhere in this volume (Macintyre et al.), and the nomenclature used therein for localities is followed throughout this discussion. METHODS Between 1991 and 1996 I made 17 visits to the Pelican Cays, during which I documented the occurrence of ascidians in all major habitats, including the mangrove ponds, all the ridges connecting the mangrove islands, and most of the reef area to the south of the islands, in addition to the two ponds at the Lagoon Cays (H & I). All observations were made by free swimming with face mask and snorkel, working slowly around the margins of the ponds, examining hanging roots and the margin of the peat bank and any other structures on which ascidians might grow. The same technique was used on the ridges and reef flats, usually with the assistance of a companion so that stones and boulders could be raised and their undersurfaces examined. All such stones were usually replaced in their original position after examination. In a few instances, scuba divers assisted by raising stones from the deeper parts of the reef or ridge or collecting specimens from the bottom of the ponds. With experience, most species of ascidian can be identified in situ in the field; in cases of doubt, specimens were collected, returned to the laboratory, relaxed with menthol for 4 to 5 hours, fixed in formaldehyde, and later transferred to 70% ethyl alcohol for storage. THE ASCIDIAN FAUNA Thus far, 117 species of shallow water ascidians in 39 genera have been recorded from the Caribbean (cf. Monniot and Monniot, 1984; Van Name, 1945; Millar, 1962; Millar & Goodbody, 1974; Goodbody, 1984a, 1984b, 1993, 1994). Specimens recorded for the ponds and reefs of the Pelican Cays total 70 species in 30 genera, thus making this locality an area of high diversity. By comparison, Monniot and Monniot (1994) have recorded 93 species in 36 genera from the island of Guadeloupe in the eastern Caribbean. Further exploration will almost certainly increase the inventory of species at both locations. 305 The occurrence and distribution of all species of ascidian recorded from the Pelican Cays and their associated reefs and ridges are summarized in Table 1. It is noteworthy that representatives of Corellidae have not been recorded from the Pelicans although two species of that family (i.e., Corella minuta, Rhodosoma turcidum) were expected to occur there. Corella minuta Traustedt, 1882, is a solitary globular species with a glassy translucent test often speckled with yellow pigment. It is normally a reef species that occurs under stones, often in clusters of four or more zooids. It has been recorded elsewhere on the barrier reef but not as yet at the Pelican Cays. Rhodosoma turcicum (Savigny, 1816) is another small solitary species found elsewhere in the Caribbean in mangrove-root communities (cf. Van Name, 1945; Millar, 1962; Goodbody, 1993). I have no records of it in the Pelican Cays or elsewhere on the barrier reef. Background information on the ascidian fauna of the Pelican Cays and on the need for management and conservation of the area has been published elsewhere (Goodbody, 1995). A full account of ascidian taxonomy is given in Kott (1985, 1992) and in the present paper the taxonomic sequence of the families in the suborders follows her arrangement. In keeping with Kott’s classification, Rhopalaea abdominalis is retained in the Diazonidae in the suborder Aplousobranchia, which differs from C. Monniot (1983a) who places this species in the Cionidae in the suborder Phlebobranchia. Suborder: Aplousobranchia Family: Polyclinidae Aplidium antillense (Gravier, 1955) There is a single unconfirmed record of this species from the ridge at the entrance to Pond A, but I have no other records of its occurrence. Elsewhere in the Caribbean it 1s recorded from Martinique and Guadeloupe (F. Monniot, 1983b) and from Port Royal Harbour, Jamaica (Goodbody, unpublished data). Aplidium bermudae (Van Name, 1902) A. bermudae is common on the peat bank surrounding some of the ponds and is especially abundant in the northwest corner of Pond C, where it grows on the top of a sloping peat bank; colonies are usually small cushions with an opalescent blue tinge to the test. The species is also frequently found on the undersurface of stones on reef ridges. Aplidium constellatum (Verrill, 1871) A single, very flattened colony was collected from the peat bank in Pond C in April 1994 and is the only record I have from the Belize Barrier Reef. Van Name (1945) considered the species to be very rare in the West Indies; F. Monniot (1983b) records it from Guadeloupe, but it was not recorded by Millar (1962) in the Caribbean collections examined by him, and no specimens were found in the extensive Caribbean collections made by P. Wagenaar Hummelinck between 1930 and 1973 (cf. Goodbody, 1984a). I have no records of the species in Jamaica. Aplidium exile Van Name, 1902 This species usually forms much larger cushion-like colonies than A. bermudae and, unlike that species, is usually pink or bright red. It is not a common species and though 306 Table 1. Occurrence of ascidian species at Pelican Cays. The ponds are identified by their respective letter or letters. R represents the reefs and ridges in the vicinity of the ponds. Taxonomic sequence follows Van Name (1945). (X = present, ? = uncertain occurrence). Locations F Species G GG H HS I Family Polyclinidae Aplidium antillense ? Aplidium bermudae xe OX x Aplidium constellatum xX Aplidium exile xX XxX Xx xX Polyclinum constellatum xX xXx X xX Aplidiopsis stellatus X Family Euherdmaniidae Euherdmania fasciculata X Trididemnum cyanophorum | X Trididemnum hians X X xX xX Xx Trididemnum orbiculatum X X X xX Trididemnum solidum xX Didemnum amethysteum X Didemnun cineraceum X Didemnum conchyliatum* x 2 2K xX X Xx Didemnum duplicatum xX X xX xX X Didemnum inauratum X Didemnum perlucidum X X Didemnum psammathodes xX X x Lissoclinum abdominale X X xX X x Lissoclinum fragile Xe EX ex xX XS XG Lissoclinum verrilli Ks xX X DOK Diplosoma glandulosum ne 2S DS xX X X X x Diplosoma listerianum* xX X X X mS I 2S X Diplosoma virens Xx Family Polycitoridae Distaplia bermudensis Ko Xx X xX Distaplia corolla » rm, a, Sa, Gam, 6 » a, 4 xX X Xx Cystodytes dellechiajei me OK xX X xX Xx Eudistoma capsulatum xX xX xX Xx Eudistoma clarum X X X X xX Eudistoma obscuratum We X xX xX xo xX Eudistoma olivaceum Xie weXoul XG EX Xs KO EX Xx xX Clavelina oblonga X Clavelina picta 2K xX ek ON 307 Table 1.--continued Locations Species RAC Bs BBs CD EEG GG EH ASI Clavelina puertosecensis* Xe Xx 1 Ok OK DS Family Pycnoclavellidae Pycnoclavella belizeana xX Family Diazonidae Rhopalaea abdominalis SK Family Perophoridae Perophora carpenteria a, < xX xX 2G Pky DK x Xx Perophora multiclathrata xX Perophora regina x Perophora viridis DEV IIE POE OR Dw ? Ecteinascidia conklini x Ecteinascidia minuta xX X X xX X XxX XX x Ecteinascidia styeloides XxX Xx xe OX Ecteinascidia turbinata* xX xX X xX Family Ascidiidae Ascidia corelloides xX Ascidia curvata xX Ascidia interrupta XE KO Ne Ket XC XG XG OX. eX x Ascidia nigra We WG 2 Mh WE WE ME WE BC OE OK x Ascidia sydneiensis xX X x xX Xx Family Styelidae Botrylloides nigrum mM Dh DK xX xX xX Botrylloides perspicuum UWE IK OE xX xX Botryllus planus X x x Botryllus tuberatus a OK Xx xX Symplegma brackenhielmi WK Ik OK xX xX X XI XS XOX Symplegma rubra X Tibitin halimedae x Polyandrocarpa tincta Ye DK xX INTENT EX: Polycarpa aurita Xe X X xX x Polycarpa cartilaginea X Polycarpa spongiabilis x eX eX me DK DK 2k OK X X Polycarpa tumida X Styela canopus mM DK OK xX WM KD X X Family Pyuridae Herdmania momus xX xX Xx X X x Pyura lignosa MS MS ML DK DS CG 2 X X Pyura munita X X X X Xx X X Pyura vittata DS Pn rk X re Ok O* X xX X 308 Table 1.--continued Locations Species R B BB CD E FG GG Hino nninN ne Microcosmus exasperatus WE OK KOK OK De OK OK DOK OK OK ox Halocynthia microspinosa xX xX Xx XE x Bathypera goreaui X Family Molgulidae Molgula occidentalis “See Plate 3 essentially a reef-dwelling species living under stones, it has been recorded from the peat bank in several of the ponds at Pelican Cays, notably Ponds A and E. Polyclinum constellatum Savigny, 1816 Elsewhere this species is sometimes abundant in mildly eutrophic lagoons, as in Lagoen Boekoeti, Aruba; Piscadeera Bay, Curagao; and Port Royal, Jamaica (Goodbody, 1984a, 1984b, 1993). In the Pelican Cays it was not common anywhere except in Pond B, where it accounted for 7% of all ascidian species recorded on mangrove roots over a 50-m stretch of pond bank in August 1995. It also occurs occasionally in other ponds, notably Ponds BB, I, and in the side arm of Pond H; on the reef ridges it seems to occur only in places where there are high levels of suspended particulates. Aplidiopsis stellatus Monniot, 1984 This species forms small greenish-yellow colonies of stalked capitate heads usually growing on the undersurface of stones on the reef. It is not common, and I have only recorded it from the ridges. Family:Euherdmaniidae Euherdmania fasciculata Monniot, 1983 The species forms small translucent colonies on the undersurface of stones on reefs and ridges. When first observed, the colony bears a superficial resemblance to a small bubble on the stone. The species is relatively common on the southern ridges. Each colony consists of three or four capitate lobes joined to one another, each of which contains two or three zooids; the whole colony is 3 to 4 cm high, and individual zooids are about 3 mm long. Family: Didemnidae The taxonomy of this large family is still poorly understood. Most species form sheet-like colonies, usually containing an abundance of small calcareous star-shaped spicules in the test substance. With few exceptions, individual species are very difficult to recognize in the field. The best contemporary guide to the identification of Caribbean species has been compiled by Francoise Monniot (1983a, 1984) describing the didemnids of Guadeloupe in the eastern 309 Caribbean. Trididemnum cyanophorum Lafargue et Duclaux, 1979 Colonies form clusters of rounded lobes, usually pinkish or purple in color owing to the presence of symbionts. It occurs on the reefs and ridges but is not common. A small colony was collected from the reef crest at Pond A in April 1993. For comments on the relationship between this species and T. solidum, see below. Trididemnum hians Monniot, 1983 The species is common in the ponds, often growing as flat sheets on the exposed peat bank. Colonies are usually translucent grey and in the field are easily confused with Cystodytes dellechiajei. In T. hians the spicules tend to aggregate toward the base of the colony around the abdominal region of the zooids; confusion can occur because C. dellechiajei also has spicules surrounding the abdomen of each zooid, albeit spicules of a quite different nature (vide infra) and in each species the test is sufficiently transparent for the spicules to be visible through it. Trididemnum orbiclatum (Van Name, 1902) The species forms white colonies but lacks any clear distinguishing features that would make it easy to recognize in the field. It has been recorded only from Ponds A, C, and H, where it was found growing on roots or directly on the exposed peat bank, and sometimes on the test of large solitary ascidians. Trididemnum solidum (Van Name, 1902) Although F. Monniot (1983a) considers 7. solidum and T. cyanophorum to be perhaps synonyms for one another, I have kept the species separate here, as is customary, on the grounds that typical 7. solidum lacks the conspicuous pink or purple coloration of 7. cyanophorum and forms larger mound-like whitish-grey colonies on reef structures. Although I have seen typical 7. cyanophorum on Pelican Cays reefs, | have seen typical 7. solidum infrequently. Didemnum (Polysyncraton) amethysteum (Van Name, 1902) A small elongate colony of this colorful species was collected from the rim of pond A in the southeast corner of the pond on April 12, 1992. This is the only record I have of this species in the Pelican Cays but, its occurrence on the pond rim suggests that it may occur more commonly in cryptic situations on reefs and ridges, and its presence may have been overlooked because of the difficulty of exploring these cryptic situations without damaging the reef structure. Didemnum cineraceum (Sluiter, 1898) Like many other didemnids, the species is difficult to identify in the field. Colonies form flat, greyish encrusting sheets, with the zooids usually showing white through the test. The species has been found only in Pond A. Didemnum conchyliatum (Sluiter, 1898) (Plate 3a) D. conchyliatum was the most common of all didemnids found in the Pelican Cays and elsewhere on the Belize Barrier Reef. In the ponds and lagoons, it grows on Rhizophora roots and directly on the exposed peat bank and is often a bright orange, sometimes a dull grey; the 310 abundance of spicules in these colonies gives them a rather rigid consistency, in comparison with the more fleshy colonies of D. duplicatum (vide infra); the grey colonies of D. conchyliatum are particularly common in Pond H at Lagoon Cays. The species is also common under stones on reefs and ridges; such colonies are usually pure white. Nevertheless positive identification in the field is not easy. Didemnum duplicatum Monniot, 1983 When growing on Rhizophora roots in the ponds, the species is relatively easy to recognize: it forms large hanging fleshy lobes with a slight pinkish tinge; this coloration is due to the reddish color of the zooids showing through the test. The species is fairly common in the ponds and in April 1991 was one of the dominant species in Pond A, but one year later, on April 12, 1992, its numbers had declined (cf. also Diplosoma listerianum). Didemnum inauratum Monniot, 1983 There are no obvious characteristics by which this species may be recognized in the field. I have collected specimens on two occasions in Pond A. Didemnum perlucidum Monniot, 1983 Several specimens identified as belonging to this species have been found in Pelican Cays. One was collected from a mangrove root in Pond A on May 10, 1992. Five other specimens attributed to this species have been collected from beneath stones on coral ridges adjacent to Manatee Cay, Co-Cat Cay, and Ridge Cay. Didemnum psammathodes (Sluiter, 1895) This is a widely distributed tropical species easily recognized by its muddy-brown coloration, which is due to an accumulation of faecal pellets in the test substance. Although it occurs elsewhere on the Belize Barrier Reef (i.e., at Twin Cays) and might be expected to occur commonly in the Pelican Cays wherever there are high levels of suspended particulates in the water, I have only recorded it from Ponds B, BB, and GG, which are mildly eutrophic (cf. Macintyre et al., this volume). Lissoclinum abdominale Monniot, 1983 This species grows as flat gelatinous, grey sheets and in the field can easily be confused with Diplosoma listerianum. Monniot (loc. cit.) described it as viscous (glaireuse) in consistency. In the Pelican Cays I have recorded it from Ponds A, C, G, and J. In Pond A it was frequently found under the lip of the peat bank on the western margin. This habitat is similar to that in which it occurs at Twin Cays further north on the barrier reef. In the type locality in Guadeloupe, F. Monniot (1983a) reports it as growing on the tunic of other ascidians, on coral, and on algae. In Jamaica I have only seen it growing on oyster shells and artificial substrates in an oyster farm. Colonies frequently have a greenish tinge owing to the presence of endosybiont algae. Lissoclinum fragile (Van Name, 1902) This is a common species in many of the ponds, usually growing on mangrove roots, on other ascidians, or on the shells of bivalve molluscs. It tends to grow as flat, snowy-white sheets, and as the name implies it is fragile, the test tearing easily to expose the zooids beneath; the Sylil abdomen of each zooid is usually orange, which further assists in identifying the animal in the field. Lissoclinum verrilli (Van Name, 1902) This is a common Caribbean species found frequently on the undersurface of stones on reefs and ridges. The colony forms a soft sheet, usually white in color owing to aggregation of spicules around the zooids. The distribution of zooids and spicules gives the colony a characteristic mottled white appearance and makes identification in the field relatively easy. The spicules are different from those of most didemnids having a tetrahedral form (cf. Van Name, 1945). In the Pelican Cays I have recorded it from most of the reef area and from the peat bank in Ponds A, C, E, and H. In the earlier literature the species is often referred to by its synonym Echinoclinum verrilli. Diplosoma glandulosum Monniot, 1983 This is a common species in most Pelican Cays ponds. It is particularly abundant at the northern end of Pond C, where it hangs down from the mangrove roots and peat bank in colorful arrays of gelatinous sheets; in other places it grows freely on the blades of turtle grass (Thalassia testudinum) on the pond floor. The coloration in this species varies greatly wherever it occurs in lagoons and ponds on the barrier reef; colonies vary from yellow to green, or mottled black and white. I have only a few records of the species from outside the ponds, on reefs and reef ridges, where it occasionally occurs under stones or on the bases of coral heads or the basal stems of octocorallians. Diplosoma listerianum (Milne-Edwards, 1841) (Plate 3b) This is one of the most common colony-forming ascidians in the Caribbean. In mangrove ponds it usually grows as grey gelatinous sheets over mangrove roots and the peat bank and over other sessile organisms in the community, often occupying the interstices between adjacent organisms. The normal coloration is translucent grey with the orange of the stomachs of zooids showing through the test. However, when we first visited Pond A in April 1991, many of the mangrove roots had large drapes of a dark green morph of this species. This coloration is quite different from that in D. virens (vide infra) and is due to pigment granules in the test substance. This seems to be similar to the color morph reported by F. Monniot (1983a) from Guadeloupe but is a form that I have only seen at the southern end of Pond C, in GG ponds, and on a single occasion on the base of an octocorallian on one of the ridges. Diplosoma virens (Hartmeyer, 1909) This is an Indo-West Pacific species for which there do not seem to be any previous authenticated records from the Caribbean. It is a reef-dwelling species and has never been seen in the ponds. It is a vivid green and forms soft, flat sheets in which the zooids are embedded. The green coloration is due to symbiotic algae (Prochloron). The species is fairly common on the reefs and ridges of the Pelican Cays, often growing on the edge rather than the top of a stone exposed to sunlight. The species is probably common throughout the barrier reef as I have other records from Tobacco Range and from the Sand Bores between Carrie Bow Cay and Wee Wee Cay. 312 Family: Polycitoridae Distaplia bermudensis Van Name, 1902 This is a common reef species usually growing as small cushion-like colonies on the undersurface of stones. It also occurs commonly on the peat bank of some of the ponds, especially Pond A and Pond C. In the latter pond it is common in the northwest corner, where it usually occurs as small brown and white colonies easily recognized by the central cloacal aperture surrounded by individual zooids, flecked with white pigment. Distaplia corolla Monniot, 1974 Originally described from the Azores, this species does seem to have been widely recognized in the Caribbean. F. Monniot (1983c) has reported on its occurrence in Guadeloupe, and I have occasionally seen it on reefs in Jamaica. Colonies are usually circular cushions with a central cloacal aperture on the upper surface, surrounded by a number of individual zooids. There are two distinct color morphs, the commonest being bright orange and less commonly a deep purple form. The species is common in all Pelican Cays ponds, usually in the orange form, and sometimes both morphs live side by side. They are equally common on mangrove roots and on the peat bank, particularly on the lip of the peat bank in Pond H in Lagoon Cays. The species is probably common throughout the barrier reef as it is an abundant component of sessile communities at Twin Cays and on pier pilings at Southwater Cay. Occasionally, D. corolla was found on the reef or reef ridges, in which case it was usually the purple morph found under an overhanging lip of rock where wave-wash may create a fairly strong current of water. Such a situation is perhaps a microcosm of what may occur on the undersurface of the peat bank lip in the ponds. Cystodytes dellechiajei (Della Valle, 1877) This is one of the most common colonial ascidians in the Pelican Cays and perhaps elsewhere in the Caribbean, although F. Monniot (1983c) does not seem to have found it to be particularly common in Guadeloupe. However, it is often one of the most difficult species to recognize in the field. Typically, colonies form rounded cushions, firm in consistency and varying greatly in color. When growing under stones on the reef, it is often predominantly white, but it also grows in great abundance on the peat bank surrounding most of the ponds as well as directly on Rhizophora roots and in these situations is more often brown, black, or a smoky grey, sometimes with a greenish tinge. Many colonies have circular depressions on the upper surface marking a common cloacal area into which the atrial apertures of systems of zooids discharge. The simplest method of identification in the field is to slice off a part of the colony with a dive knife. This reveals the dense accumulation of calcareous discs that surround the abdominal region of each zooid. These discs form a small cup in which the abdomen sits, and they are readily recognizable with the naked eye. However, one must not confuse this situation with the layer of dense spicules (tiny star-shaped calcareous inclusions) found in species such as Trididemnum hians (see above). Eudistoma capsulatum (Van Name, 1902) Several specimens considered to belong to this species have been collected from the peat bank of some of the ponds (e.g., Ponds A, C, and E). 313 Eudistoma clarum (Van Name, 1902) This species forms gelatinous cushions of a glassy appearance. Often the colony is globular in form. It is relatively common along the peat bank and on mangrove roots in Ponds A and H, but I have not found it in any abundance in other Pelican Cays ponds. Eudistoma obscuratum (Van Name, 1902) Like other species in the genus, E. obscuratum forms rounded cushion colonies but is usually small (1 to 2 cm in diameter) and black. It is fairly common on the peat bank in many of the ponds and occasionally occurs on the reef, often in rock crevices in shallow water, where it is fully exposed to bright illumination. In Pond H, colonies are sometimes large rounded cushions as much as 13 cm in diameter. Eudistoma olivaceum (Van Name, 1902) The species has a variable colony form, depending on where it is growing. Small colonies form rounded cushions with a central cloacal aperture surrounded by individual zooids but in quiet lagoons large colonies may develop in which each group of zooids is supported on a long stalk, and clusters of these stalked structures grow closely together on mangrove roots or other supports. Colonies are usually greenish yellow but large stalked clusters may sometimes have a reddish tinge, particularly on the supporting stalk. Although FE. olivaceum is abundant elsewhere on the barrier reef, especially in parts of Twin Cays, it is relatively rare in the Pelican Cays. I have recorded it in small numbers from most of the ponds (see Table 1), and in such instances the colonies have been small and isolated. Clavelina oblonga Herdman, 1880 This is probably the best known of all Clavelina species in the Western North Atlantic region (see Van Name 1945; F. Monniot, 1983c). It forms clusters of glassy translucent zooids united by vascular structures at their posterior ends and by fragile union of the test substance of adjacent zooids. White pigment flecks may occur in individual zooids. Elsewhere in the Caribbean it sometimes forms massive colonies on mangrove roots (Goodbody, 1993), but I have never seen it occupy such a habitat in the Pelican Cays; instead it sometimes forms small colonies under stones on the reef and reef ridges. Van Name (loc. cit.) records it from similar situations, and at Point Gourda in Trinidad I have recorded it growing on reef structures in about 2 m of water. Clavelina picta (Verrill, 1900) (Plate 3c) Morphologically, the species differs little from C. oblonga (see Van Name, loc. cit.) but is readily distinguished by its pinkish color. It is common throughout the ponds of Pelican Cays, growing on mangrove roots and sometimes directly on the peat bank or often on the edge of the peat lip, where it is well supported clear of any sediments. It frequently grows in close association with C. puertosecensis (vide infra) in such a manner as to make it difficult to distinguish one colony from the other except by coloration. Clavelina puertosecensis Millar & Goodbody, 1974 (Plate 3d) The species was first described from very small colonies collected on the fore reef at Discovery Bay, Jamaica, in about 60 m of water; it has subsequently been recorded from a reef in 314 Guadeloupe by F. Monniot, (1983c). In the Pelican Cays large colonies are abundant in many of the ponds and are immediately recognizable by their deep blue color as opposed to the pink coloration of C. picta. Like the latter species, C. puertosecensis commonly grows on mangrove roots and on the lip of the peat bank. Both species also often grow on the bases of octocorallians on the reef ridges, especially the ridge extending south from Co-Cat Cay. In Bermuda and on the South Shelf of Jamaica, C. Picta is also found growing on octocorallians. Interestingly, although both of these species are abundant in Ponds A and C, they seem to be totally absent from Pond H, where solitary species of ascidian prevail over colonial forms. Neither species occurs at Twin Cays farther north on the barrier reef, but C. Picta was found to be common in one lagoon in Blue Ground Range. Family: Pycnoclavellidae Pycnoclavella belizeana Goodbody, 1996 This species was first described from specimens collected in the peat bank of one of the mangrove channels at Twin Cays 12 km north of the Pelican Cays. It is a colony-forming species with tiny (1-mm-long) erect zooids arising from vascular structures growing through the surface of the peat at the edge of the mangrove bank. In the Pelican Cays it has been recorded only from the southern shore of Pond G, where it grows in a similar peat habitat. Zooids of colonies seen or collected at Twin Cays are all pure white, while those from Pond G are yellowish. The wide spatial separation of the two localities at which the species has been recorded suggest that it may be widely distributed throughout the barrier reef. The yellow morph seen at Pond G is easily recognizable in the field, but the white morph from Twin Cays is very difficult to detect, particularly as the zooids retract into the substrate if in any way disturbed. Family: Diazonidae Rhopalaea abdominalis (Sluiter, 1898) The normal habitat of this species seems to be in relatively deep water on coral reefs (personal observation; see also Van Name, 1945). It is surprising, therefore, to find that it is relatively common in shallow water in Pond A at Pelican Cays, where it grows on hanging mangrove roots and directly on the peat bank. It is a large maroon ascidian with a thick test containing one or two small zooids. I do not have records of it from any of the other ponds, but several specimens have been collected by scuba divers on the reef ridges at 5 to 10 m. Suborder: Phlebobranchia Family: Perophoridae Two genera of this family are common in the Caribbean, Perophora and Ecteinascidia. For a review of the genus Perophora in the Western Atlantic, see Goodbody (1994). Of the five species reported in that paper, only four have been recorded in the Pelican Cays. 315 Perophora carpenteria Goodbody, 1994 Perophora carpenteria resembles other species in the genus by having a colony of tiny globular green zooids connected together by creeping stolons. It is not possible to identify this species in the field, but it is the commonest species of Perophora on the pond rims, where it creeps along the peat bank and often grows on the blades of seagrass (Thalassia). Perophora multiclathrata (Sluiter, 1904) The species is difficult to recognize in the field and although its presence in the ponds may have been overlooked, I have only collected specimens from the reef area. Perophora regina Goodbody & Cole, 1987 This species is easily identified by its large zooids, which form bushy colonies on mangrove roots. Each zooid has a conspicuous white spot bordered by yellow pigment between the two siphons. P. regina has been recorded only from Pond B, where it is sometimes found in profusion on mangrove roots. This is the only place outside the type locality at Twin Cays where I have recorded this species in abundance, although some very small colonies have been found in shallow lagoons in Blue Ground Range just south of Twin Cays. Further exploration is likely to reveal a wide distribution in lagoons on the barrier reef. Perophora viridis Verrill, 1871 P. viridis was found in many of the ponds growing among other members of the sessile community; however, it is not common. Positive identification in the field is difficult. Ecteinascidia conklini Berrill, 1932 The species was originally described from Bermuda, but its occurrence and distribution throughout the Caribbean have since remained in doubt owing in part to confusion with other species of Ecteinascidia (cf. C. Monniot, 1973, 1983a). The species is rare in the Pelican Cays but has occasionally been found under stones on reef ridges or on the basal stems of octocorallians. It is abundant in one of the small lagoons at Blue Ground Range farther north on the barrier reef. The species is readily recognizable by its bright yellow-green coloration and a simple ring of red pigment around each of the siphons. There is an unconfirmed report of a very small colony on a mangrove root on the ridge at the entrance to Pond A in Pelican Cays. Ecteinascidia minuta (Berrill, 1932) This is the smallest of all Caribbean species in the genus. Individual zooids are flattened and attached to the substrate by the ventral surface, the entire colony forming a flattened group of pale green zooids attached to one another by vascular stolons. It is fairly common under stones on reef ridges and occasionally grows on mangrove roots or other hard substrata such as oyster shells. I have not seen it growing directly on the peat bank. Ecteinascidia styeloides (Yraustedt, 1882) The species forms clusters of translucent green zooids intermediate in size between those of E. turbinata and E. conklini. Many records of E. conklini in the literature may indeed refer to this poorly recognized species (cf. C. Monniot, 1983a). E. styeloides is abundant in mangrove channels at Twin Cays farther north on the barrier reef but has only been found at a few locations 316 in the Pelican Cays (cf. Table 1). The records of the species in Ponds B and I and the side arm of Pond H are not surprising as these ponds provide the type of food-rich environment in which this species thrives. Specimens have also been recorded on the ridges. Ecteinascidia turbinata Herdman, 1881 (Plate 3e) ; This is the most easily recognized of all Caribbean species of Ecteinascidia. It forms clusters of bright orange zooids connected at their posterior ends by a mass of vascular structures. Although it is widely distributed in mangrove lagoons throughout the region (cf. Van Name, 1945: Goodbody, 1984a, 1984b, 1993), it is not common in Pelican Cays; isolated colonies have been found around the perimeter of Pond A, particularly at the northern and eastern shores. The species is quite abundant in a small isolated pond (BB) on the southwestern side of Manatee Cay, and small colonies occur at the south end of Pond C and in Pond GG. Elsewhere on the barrier reef it is probably fairly common and I have recorded it from mangrove roots in channels at San Pedro and farther north on the reef, but not at Twin Cays. Family: Ascidiidae All of the species in this family are solitary forms, sometimes quite large, up to 25 cm in length. Ascidia corelloides (Van Name, 1924) This is a fairly cryptic species, usually found under stones on reefs and ridges. It is small and grey-green. Although it is common elsewhere on the barrier reef (e.g., Carrie Bow Cay), it is not common on Pelican Cays ridges or reefs and has never been seen in the ponds. Ascidia curvata (Traustedt, 1882) The species seems to be a common inhabitant of reef structures in the Caribbean, often growing on the underside of stones. It is only in such habitats that I have found it in Pelican Cays, although at Port Royal in Jamaica it has recently invaded mangrove ponds and grows commonly on Rhizophora roots. C. Monniot (1983a) considers this species name a nomen conservandum and lists various synonyms; he further describes a new species, Ascidia tenue, suggesting that this also is a synonym (nom nouveau pour Ascidia curvata). The nomenclature is therefore confusing, and it seems appropriate to retain Traustedt's original name as a nomen conservandum. Ascidia interrupta Heller, 1878 This is one of the commonest species of ascidian in Pelican Cays. It is grey-green and often 25 cm or more in length. It is common in most of the ponds, often attached to bottom rubble or to mangrove roots, but often partly buried in the peat bank, a niche also occupied by the species in Jamaica (Goodbody, 1966, 1993). At Pond A, I found large populations living among the plates of Agaricia tenuifolia on the inner face of the ridge that crosses the mouth of the pond (cf. Macintyre et al., this volume). Ascidia nigra (Savigny, 1816) This is an abundant species throughout the Pelican Cays ponds. It is large, 25 cm or more, and jet black in color, so is quite unmistakable in the field. Unlike A. interrupta, which often SLY) lives close to the sediments or embedded in the peat (vide supra), A. nigra seems to prefer living raised well above the substratum and is thus commonly found on mangrove roots and on the lip of the peat bank. It is common among the plates of Agaricia tenuifolia at the entrance to Pond A, and in Pond C is common on the shell gravel on the western rim. In this latter case the shells appear to provide sufficient support to keep the ascidians well clear of the detrital sediments. This preference for a site raised off the sediments has also been recorded in Jamaican communities (Goodbody, 1966). Pond H at Lagoon Cays has enormous populations of Ascidia nigra, mixed with other solitary species, surrounding the entire rim and covering the lip of the peat bank and its undersurface. Pond H also has relatively few colonial species, except Distaplia corolla. The abundance of solitary species such as 4. nigra and Microcosmus exasperatus is not easily explained. (Ascidia nigra is a synonym for Phallusia nigra, but A. nigra is in common usage in the Caribbean. For a discussion of these two genera, see Van Name, 1945; and Kott, 1985.) Ascidia sydneiensis Stimpson, 1855 This is a large species, probably often 30 cm in length, but it is difficult to obtain intact specimens. It lives on the bottom, often embedded in sediments, and the only visible sign of its presence is the large bright yellow-green branchial siphon protruding from the sediment. It seems to be fairly common throughout the Pelican Cays, usually at the bottom of ponds and sometimes along the margins of reef ridges, which is where divers have collected specimens attached to the underside of stones. For instance, Ian and Eve Macintyre recovered a specimen from a depth of 10 m on the east side of the ridge between Fisherman's Cay and Ridge Cay on July 25, 1995. A. sydneiensis is fairly common around the rim of Pond A in bottom sediments, especially on the eastern side. Suborder: Stolidobranchia Family: Styelidae Botrylloides nigrum Herdman, 1886 This is a common Caribbean species often abundant in mangrove root communities. It is usually brick red or orange and forms flattened sheets. Although it has been found in several of the ponds at Pelican Cays, it is not abundant at any of these locations. Herdman's original specimens from Bermuda appear to have been black in color, and although I have seen such color morphs at Bonaire in the Netherlands Antilles (Goodbody, 1984b), I have not seen them elsewhere in the Caribbean. Black forms of the species were recorded from Pond B in Pelican Cays in August 1995. The species is an uncommon inhabitant of the reefs and ridges. Botrylloides perspicuum Herdman, 1886 This is an Indo-Pacific species whose presence in the Caribbean has not previously been confirmed. Like B. nigrum, it has a flattened growth form, usually thicker than the latter species, and with conspicuous areas of vascularized test between paired rows of zooids (cf. Kott, 1985). It is frequently green or pinkish in coloration. The species appears to be common on the Belizean barrier reef as I have recorded it in several ponds in Pelican Cays as well as at Twin Cays and South Water Cay. Botrylloides perspicuum was first described from specimens collected by the 318 Challenger Expedition in 10 fathoms (18 m) of water off the Philippine Islands. Hartmeyer (1909-11) includes this species in a list of species occurring at Bermuda but provides no details on the precise location; Van Name (1945) dismisses this report as being "evidently a mistake." Botryllus planus (Van Name, 1902) The species is small and inconspicuous, often maroon in coloration. It inhabits mainly reef structures but has occasionally been found growing on mangrove roots around pond rims. Botryllus tuberatus Ritter and Forsyth, 1917 This small species is often black and white, pink, or greenish in color. It is fairly common on ridges or on reefs, where it seems to prefer flat surfaces such as dead plates of Millepora. It has been recorded from several ponds, where it often occupies space on dead mangrove leaves lying on the bank or pond bottom. Symplegma brackenhielmi (Michaelsen, 1904) A colonial species that forms flat sheets with a characteristic mottled appearance of black, green, and cream. It is common in many of the ponds, where it grows either directly on the peat bank or on large solitary ascidians or oysters. Small colonies having only a few dark green zooids, each with a yellow ring of pigment in the vicinity of the siphons, may also belong to this species, or they may be colonies of Symplegma viride, but as yet I have not been able to confirm this identification. Such colonies are found frequently under stones on the reef and occasionally in the ponds on the peat bank or on leaf litter. Symplegma rubra Monniot, 1972 Colonies are encrusting, usually growing directly over mangrove roots or other organisms such as solitary ascidians and oysters. Individual zooids appear slightly bulbous, colored either bright yellow, pink, or red. Although it is common elsewhere in the Caribbean—for example, Guadeloupe (C. Monniot, 1983b) and Jamaica (Goodbody, 1993)—it appears to be rare on the Belizean barrier reef. In Pelican Cays I have recorded it on a single occasion in Pond E in April 1994. It has not been seen in any of the other ponds, nor do I have records of it at Twin Cays or Blue Ground Range farther north on the barrier reef. Tibitin halimedae Monniot, 1983 The species forms creeping colonies of tiny globular zooids, initially resembling a very small Perophora. In the type locality in Guadeloupe, C. Monniot (1983b) found it growing on stems and leaves of Halimeda. It is fairly common on reefs and ridges in Pelican Cays but has always been found underneath pieces of coral rubble away from any source of fine sediments. It has never been seen in any of the ponds. In Jamaica it has been seen only on the undersurface of coral rubble (personal observation). Polyandrocarpa tincta Van Name, 1902 The species is common throughout the Pelican Cays. In ponds it often forms thick leathery colonies, deep maroon in color, growing as a sheet over a mangrove root or covering other organisms. The individual zooids are bulbous. On the reefs and ridges, it was found as smaller colonies under coral rubble, usually with bright red zooids, often widely spaced. a9 Polycarpa aurita (Sluiter, 1890) This species is common throughout the ponds of Pelican Cays. It is easily distinguished in the field from Polycarpa spongiabilis by its erect cylindrical form and by its siphons, which have a maroon rim and creamy interior. The atrial siphon is far back on the dorsal side. Like P. spongiabilis, it is frequently found on the peat bank and roots and among Thalassia on the pond bottom. The species does not seem to have been widely recognized in the Caribbean. I have no records of its occurrence in Jamaican waters, nor have | recorded it farther north on the Belize barrier reef at Twin Cays or surrounding areas. Van Name (1945), using the synonym Polycarpa circumarata (Sluiter, 1904), reports on its occurrence in the Gulf of Mexico, Panama, Cura¢ao, and La Tortuga Island north of Venezuela. The species also seems to be widely distributed in Australia and other parts of the Indo-Pacific region (Kott, 1990). The species has been recorded under a number of different names; for a synonymy see Van Name (loc. cit.). Polycarpa cartilaginea Sluiter, 1898 This is a small species, seldom exceeding 1.5 to 2.0 cm in length, usually a dull grey- green. It is frequently found on the underside of stones and coral rubble on reefs and ridges, but I have never found it in the Rhizophora root communities in the ponds. In August 1991, three specimens were collected from the ridge guarding the entrance to Pond A; one of these was attached to the test surface of a specimen of Polycarpa spongiabilis. Polycarpa spongiabilis Traustedt, 1883 This is one of the most common solitary ascidians in the ponds; zooids are usually squat and bulbous, a dirty grey-green in color. When the large branchial siphon is wide open, it is possible to see the snowy white ring of tentacles just inside the aperture. The zooids are usually fairly soft in consistency. The species is common in all of the ponds, often found on bottom sediments among Thalassia, but it also grows on the peat bank and on Rhizophora roots. A different morph found in some places, especially Pond G, is pale yellow-grey with faint red lines running vertically down the test. These zooids are firm in consistency with a more cartilaginous test than the other morph. Polycarpa tumida Heller, 1878 A single zooid considered to be the solitary form of this species was collected from the ridge at the entrance to Pond A in July 1991. Styela canopus Savigny, 1816 This small styelid is common throughout the Caribbean, particularly in mangrove root communities. The species is frequently reported on by its synonym Styela partita; zooids are solitary, but as a result of larval aggregation at settlement, the adults are often found in dense clusters. In the field they can be recognized by prominent dark and light vertical stripes in the interior rim of the siphons. The test is thin and body color is an indistinct grey-green. Although difficult to see among dense clusters of other sessile organisms, the species is fairly common in the ponds but seldom seen on the reef or reef ridges. 320 Family: Pyuridae Herdmania momus (Savigny, 1816) A large solitary ascidian with a pink tinge to the test, the branchial siphon usually with a prominent iridescent green and pink lining. The body is often inflated and globular. The species is very common in Pelican Cays, especially in the ponds, where it is common on soft sediments and among 7halassia and less common on the peat bank and mangrove roots. The species is pan- tropical and common throughout the Caribbean. Pyura lignosa Michaelsen, 1908 This is a very large solitary species commonly found around the rim of most ponds in the Pelican Cays. Zooids are often deep red but not always attached to the base of a vertically growing mangrove root emerging from the peat. They may also be found among other organisms in the sessile community and growing directly on the peat bank. Large zooids may exceed 15 cm in length. The test is hard and inflexible, and close inspection will usually reveal the presence of scale-like polygonal markings on the exterior surface. Pyura munita (Van Name, 1902) A small species, usually only 1.0 to 5.0 cm in length, it occurs in bottom sediments in some of the ponds and occasionally under coral rubble on reef and ridges. Ian and Eve Macintyre collected about 60 specimens at depths of about 10 m just inside the entrance to Pond C on July 28, 1995. At some locations, as in the southwest corner of Pond E, it lives buried in the peat bank, and the only signs of its presence are the elongate siphons protruding from the peat surface; however, small individuals of Mo/gula occidentalis (vide infra) occupy a similar habitat in Pond E. It is difficult to distinguish the two in the field, although the siphons of P. munita have an iridescent bluish green lining that may be visible in good light. Pyura vittata (Stimpson, 1852) A solitary ascidian reaching a length of 9.0 to 10 cm. The test is firm and thick and usually a drab red-brown. There are no other characters by which the species is easily recognized in the field. The species is common in most of the ponds, living on Rhizophora roots and directly on the peat bank. It was also common among coral rubble on the reefs and ridges. Microcosmus exasperatus Heller, 1878 (Plate 3f) This is another rather nondescript solitary species that is difficult to identify in the field and sometimes difficult to distinguish from Pyura vittata. It occurs in two color morphs, one bright orange, the other a dull grey-green. It is common throughout Pelican Cays in both ponds and on reefs and ridges. It is particularly abundant in Pond H and is one of the most common solitary species on roots and peat banks in Ponds A, C, and E. Halocynthia microspinosa (Van Name, 1921) Van Name (1921) originally described this species under the name Tethyum microspinosum on the basis of a single specimen believed to have been collected at Andros Island in the Bahamas. Later (Van Name, 1945), he concluded that, since no other specimens had been recorded from elsewhere in the West Indies, a mistake must have been made and the 321 specimen from the Bahamas was an abnormal specimen of Halocynthia pyriformis. Millar and Goodbody (1974) reported on several specimens collected from reefs in Jamaica at depths of 3 to 60 m; they concluded that Halocynthia microspinosa is a valid species. On the basis of these specimens these authors redescribed the species. Since that time I have recorded many specimens from shallow-water reef habitats on the south coast of Jamaica, usually at depths of 0 to 10 m under pieces of coral rubble. Claude Monniot (1983c) describes specimens from Guadeloupe also found growing among coral but does not mention any records from mangrove habitats. In the Pelican Cays I have found it frequently in reef and ridge communities, usually under coral rubble. It has also been found growing in Ponds A, E, G, and H, where it occurs on Rhizophora roots and in crevices on the peat bank. The species is solitary and easily recognized in the field. Zooids are usually 1 to 5 cm long, with a red or orange-red test. The test surface is covered by tiny branched spines and there is a conspicuous ring of larger spines around the margins of the siphons. Since the species seems to be predominantly a reef species it especially interesting to find it living in mangrove ponds in the Pelican Cays, as is also the case for the normally reef-dwelling Rhopalaea abdominalis (vide supra). Bathypera goreaui Millar and Goodbody, 1974 Species of the genus Bathypera are normally found in very deep water (see Van Name, 1945). B. goreaui is an exception: the type specimen was collected on the fore reef at Discovery Bay, Jamaica, in 53 m of water, while other specimens were collected in the same area at depths of 55 to 90 m. A single specimen was collected on the ridge at the entrance to Pond A at Pelican Cays on July 31, 1991. The recorded depth was less than 2 m. Although I have no other records of the species on the Belize Barrier Reef, its occurrence in shallow water at Cat Cay suggests that it may be widely distributed in deeper parts of the reef. Family: Molgulidae Molgula occidentalis Traustedt, 1883 A solitary and inconspicuous species growing 1.0 to 3.0 cm in length. Large specimens are sometimes found among other components of the sessile community in mangrove lagoons. Small specimens sometimes occur in coarse bottom sediments, often among seagrass (Thalassia testudinum). Small specimens are also frequently found embedded in the peat bank, where the only visible sign of their presence is the two protruding siphons. In this latter respect it resembles Pyura munita and the two species are often impossible to separate in the field (vide supra). Both species are common in the peat bank in the southwest corner of Pond E. DISCUSSION As the contributions in this volume make clear, the Pelican Cays have a rich diversity of fauna and flora; this is especially true in the case of the ascidian fauna, which are far richer than those found elsewhere on the barrier reef. They are particularly abundant on the inner (south) face of the ridge at the mouth of Pond A, where one finds large populations of solitary ascidians, dominated by Ascidia interrupta, A. nigra, Microcosmus exasperatus (Plate 3f), and Polycarpa 322 spongiabilis. The ascidians grow in among the plates of the coral Agaricia tenuifolia, often partly overgrown by Zooanthus in such a manner that only the siphons of the ascidians can be seen emerging through the zooanthid cover. Apart from the solitary ascidians, this inner face of the ridge supports a diverse assemblage of colonial species living cryptically among the coral plates. However, this cryptic fauna is difficult to study without doing unnecessary damage to the reef structure. Solitary species of ascidian are also abundant in Pond H in the Lagoon Cays, but in this case they are concentrated along the western rim of the pond on the peat bank and among mangrove roots. The dominant species here are Ascidia nigra and Microcosmus exasperatus (Plate 3f), but there are also large populations of the colonial ascidian Distaplia corolla. The western rim of Pond A and the northwestern corner of Pond C are also areas of high diversity; in both cases the brightly colored colonies of Clavelina picta (Plate 3c) and C. puertosecensis (Plate 3d) are prominent features of the communities as well as many other colonial species. On the rim of Pond A, Cystodytes dellechiajei is particularly abundant. In Pond C, the western rim is the habitat of abundant Distaplia corolla, while the northern end has unusually large numbers of the colonial species Diplosoma glandulosum. Interestingly, the rims of many of the ponds are inhabited by species that otherwise are characteristic of reef environments. Among these are Aplidium bermudae, A. exile, Cystodytes dellechiajei, and Eudistoma clarum. Rhopalaea abdominalis, which normally occurs below 20 m in reef habitats, is relatively common along parts of the western rim of Pond A, and Halocynthia microspinosa, an otherwise exclusively reef species, occurs on the rim of several of the ponds but always as isolated zooids occurring only here and here. It is not clear why so many reef-dwelling species have invaded the mangrove environment in the Pelican Cays. Note, however, that in every case reef environments—such as the ridge extending across the entrance to Pond A—are in close proximity to the entrances of the ponds (for further details, see Macintyre et al., this volume). Another interesting feature is the complete absence of Clavelina picta (Plate 3c) and C. puertosecensis (Plate 3d) from Pond H in the Lagoon Cays, species that are abundant in Ponds A and C. All three of these ponds have an entrance toward the southern end and a closed northern end facing toward the prevailing wind; the mangrove community at these northern limits is sufficiently thin as to permit ingress of water from outside into the lagoon when the wind is blowing strongly. These physiographic features may have no bearing on the absence of the Clavelina species from Pond H, which might instead have more subtle associations with the abundance of solitary species in this pond. It should be noted, however, that both species of Clavelina are present on reef structures (e.g., the base of coral heads or the base of octocorallian stems) not very far from the entrance to Pond H. Hence population reservoirs of these species do occur in the vicinity. Nowhere in the Pelican Cays have I recorded colonies of Perophora bermudensis. This species is characteristic of localities having a strong water current (Goodbody, 1994), and its absence from Pelican Cays may be due to the absence of such conditions in any of the ponds. The species is common at Twin Cays 12 km further north on the barrier reef. As Table | indicates, most species of ascidian have a wide distribution throughout the various ponds and on the reef ridges; however, the composition varies from pond to pond. Ina future paper, I expect to demonstrate that only a few species account for a high percentage of total species abundance. In some cases, a single species (Ascidia nigra in the case of Ponds A and H) may contribute as much as 50% of the population and hence most of the biomass. 323 The high diversity of species along the pond rims, not only of ascidians but of sessile organisms in general (cf. other contributions to this volume) inevitably leads to intense competition between species. This competition can be seen in the frequent overgrowth of ascidians by sponges and also of one ascidian species by another. Cystodytes dellechiajei seems to be particularly aggressive in this respect, and it is common to see encounters between two colonies of this species in which one colony is overrolling the other, or Cystodytes is growing over another species of colonial ascidian. The expression "overrolling" is apt because the margins of the Cystodytes colonies are usually rounded, and at the point of overgrowth one colony appears to be rolling over the other. Similar types of encounter are frequently observed between adjacent colonies of species of Didemnum. Species of Didemnum, Trididemnum, and Lissoclinum frequently grow over zooids of solitary species such as Polycarpa, Microcosmus, and Pyura; this is not a competitive interaction but is a case of the use of secondary space for growth and development. Several species of sponges compete with ascidians for space and food supply and often grow over ascidians. By contrast, ascidian overgrowths on sponges are relatively rare and usually involve small species such as Perophora carpenteria, whose stolonic growth enables it to run over the sponge like a vine. Despite the diversity and abundance of sessile organisms along the rims of the ponds, a great deal of unused or underutilized space still exists, both at the margin of the peat bank itself and on the many bare patches on root structures. It is difficult to explain this phenomenon, for throughout the year there must be an abundance of larval forms seeking a substrate on which to settle. With mobile sediments along the peat bank constantly falling into the water below (cf. Macintyre et al., this volume), it may be that larvae attempting to settle are inhibited or dislodged by the moving sediment. Any colony-forming species that does gain a foothold may be able to spread along the bank quickly, giving the appearance of a very successful colonist. Many of the peat banks are heavily populated by the alga Lobophora variegata, and further research might show that this alga exudes inhibitory substances in its vicinity. The frequent bare patches on bank roots or hanging roots are no doubt due to quite different factors. They may represent areas in which substantial growths of sessile communities have developed and were later sloughed off under their own weight (cf. Goodbody, 1965), or surrounding communities of sponges may have inhibited settlement of other organisms (cf. Goodbody, 1961). Or they may represent areas grazed by fishes or other organisms. All the ponds have an abundance of juvenile parrot fish, which graze on the roots and peat bank, and their activities, mostly in search of algae, must adversely affect small newly settled ascidians or other sessile forms; over the long term, this could reduce the opportunity for full development of the sessile community. Other predators on the sessile communities include French Angel Fish (Pomacanthus paru), which eat sponges but are unlikely to leave a completely bare surface during their grazing; once an angel fish has finished grazing, however, other grazers might "polish" the surface. Few other forces at work are likely to have caused these bare patches. Although mangrove roots grow very rapidly, and many fresh root tips are thus free of colonists, the upper portions of the root also have unused space, which is not so new that there has not been time for sessile organisms to colonize. The phenomenon must be due to either inhibition or destruction. Other destructive forces may include grazing gastropods or flatworms, both of which might consume newly settled organisms and maintain clean space notwithstanding any attempt at sessile community development. Mature communities may be dislodged either by sloughing 324 under their own weight (vide supra) or physical disturbance as a result of wave action. Excessive wave action, particularly when created by the wash of a passing boat, causes hanging roots to bang against one another with resultant weakening of the sessile community; sloughing may be enhanced by the activities of crabs (i.e., Menippe nodifrons) burrowing among the sessile organisms. Although I have seen this happen in Jamaican mangroves, I have never seen M. nodifrons in the Pelican Cays ponds. Trunk fishes (Lactophrys) regularly attack large solitary ascidians such as Ascidia nigra. They slice open the test and eat out the body from within; attacks such as these usually leave evidence behind in the form of broken test material, but none of the bare spaces that I have examined show evidence of this kind. The Pelican Cays ponds provide an opportunity to study not only spatial differences in ascidian communities but also temporal changes. I had originally intended to carry out such a study using transect methods, but circumstances have prevented me from returning to the site to complete this work. ACKNOWLEDGMENTS Special thanks are extended to Klaus Riitzler, Director of the Caribbean Coral Reef Ecosystem Program (CCRE) at the National Museum of Natural History, for his continued encouragement and support of my research in Belize. Thanks also to Michael R. Carpenter for his assistance and companionship during much of the fieldwork, and to Bruno Pernet and Tony Rath, who also provided field assistance. I am also grateful to lan and Eve Macintyre, who collected specimens for me during scuba dives. Paul Shave of Wee Wee Cay first introduced me to the ascidian fauna at the Pelican Cays and provided logistical support on several occasions. Shakira Azan, Antroy Ashton, and Jahsen Levy provided assistance in the laboratory. I especially thank my wife, Charlotte Goodbody, who helped me compile this paper. Field work for this survey was supported by the National Museum of Natural History's Caribbean Coral Reef Ecosystem program (CCRE Contribution No.590). REFERENCES Goodbody, I. 1961. Inhibition of development of a marine sessile community. Nature 190:282-283. 1965. The biology of Ascidia nigra (Savigny) III. The annual pattern of colonisation. Biol. Bull. 129:128-133. 1966. Some aspects of the biology of the genus Ascidia in Jamaica. Proceedings of the 7th Meeting of the Association of Island Marine Laboratories of the Caribbean, Barbados, p. 3 (Abstract). 1984a. Ascidians from Caribbean shallow water localities. Studies on the Fauna of Curagao and Other Caribbean Islands 67:62—76. 1984b. The ascidian fauna of two contrasting lagoons in the Netherlands Antilles: Piscadera Baai, Curagao, and the Lac of Bonaire. Studies on the Fauna of Curagao and Other Caribbean Islands 67:21\-61. 1993. The ascidian fauna of a Jamaican lagoon: Thirty years of change. Rev. Biol. Trop., Supplemento 41(1):35-38. 325 1994. The tropical western Atlantic Perophoridae (Ascidiacea) I. The genus Perophora. Bull. Mar. Sci. 55:176-192. 1995. Ascidian communities in southern Belize—a problem in diversity and conservation: Aquatic Conservation. Marine and Freshwater Ecosystems 5:355-358. Hartmeyer, R. 1909-1911. Ascidien [continuation of work by Seeliger]. In Klassen und Ordnungen des Tier- reichs, vol. 3, edited by H. G. Bronn, 1281-1773. Leipzig. Kott, P. 1974. The evolution and distribution of Australian tropical Ascidiacea. Second International Coral Reef Symposium. Brisbane, October 1974, pp. 406-423. 1985. The Australian Ascidiacea Part 1, Phlebobranchia and Stolidobranchia. Memoirs of the Queensland Museum 23:1-440. 1990. The Australian Ascidiacea, Phlebobranchia and Stolidobranchia, Supplement. Memoirs of the Queensland Museum 29(1):267-298. 1992. The Australian Ascidiacea Part 3, Aplousobranchia (2). Memoirs of the Queensland Museum 32 (2):375-655. Millar, R. H. 1962. Some ascidians from the Caribbean. Studies on the Fauna of Curacao and Other Caribbean Islands 59:61—77. Millar, R. H., and I. Goodbody 1974. New species of ascidians from the West Indies. Studies on the Fauna of Curagao and Other Caribbean Islands 45:142-161. Monniot, C. S73: Ascidies phlebobranches des Bermudes. Bull Mus. natn. Hist. nat., Paris 3e ser No. 82, Zool. 61:939-948. 1983a. Ascidies littorales de Guadeloupe II. Phlebobranches. Bull. Mus. natn. Hist. Nat., Paris 4e ser., 5, section A, No. 1, pp. 51-71. 1983b. Ascidies littorales de Guadeloupe IV Styelidae. Bull, Mus. natn. Hist. nat., Paris 4e Ser. 5.section A, No. 2, pp. 423-456. 1983c. Ascidies littorales de Guadeloupe VI: Pyuridae et Molgulidae. Bull. Mus. natn. Hist. nat., Paris 4e ser., 5, section A, No. 4, pp. 1021-1044. Monniot, F. 1983a. Ascidies littorales de Guadeloupe: I. Didemnidae. Bull. Mus. natn. Hist. nat., Paris 4e ser., 5, section A (1), pp. 5-49. 1983b. Ascidies littorales de Guadeloupe: II]. Polyclinidae. Bull. Mus. natn. Hist. nat., Paris ser., 5, section A, No., 2 pp. 413-422. 1983c. Ascidies littorales de Guadeloupe V. Polycitoridae. Bull. Mus. natn. Hist. nat., Paris 4e ser., 5, section A, No. 4, pp. 999-1019. 1984. Ascidies littorales de Guadeloupe VII. Questions de systematique evolutive posees par les Didemnidae. Bull Mus. natn. Hist. nat., Paris 4e ser., 6, section A, No. 4, pp. 885-905. 326 Monniot, C., and F. Monniot. 1984. Ascidies littorales de Guadeloupe VII. Especees nouvelles et complementaires a l'inventaire. Bull. Mus. natn. Hist. nat., Paris 4e ser., 6 section A, No. 3, pp. 567-582. Sluiter, C. P. ; 1904. Die Tunicaten der Siboga-Expedition. Part 1. Die socialen und Holosomen Ascidien. Siboga Exped. 56a:1—126. Van Name, W. G. NOD le Ascidians of the West Indian region and south eastern United States. Bull. Am. Mus. Nat. Hist. 44:283-294. 1945. The North and South American Ascidians. Bull. Am. Mus. Nat. Hist. 84:1-476. PLATES 28 Plate 1. Marine habitats at the Pelican Cays: a, Cat Cay, red mangrove (Rhizophora mangle) anchored on reef composed mainly of lettuce coral Agaricia tenuifolia (photo, Chip Clark); b, close-up of a showing lettuce coral associated with zoanthids, mangrove stilt roots covered by sponges, and schooling fishes; c, outer slope of coral ridge, Cat Cay, composed of lettuce coral and staghorn coral (Acropora cervicornis); d, inner slope of coral ridge, Cat Cay, where lettuce coral is densely populated by sponges and crinoid echinoderms; e, red- mangrove prop roots lining pond E, Fisherman’s Cay, covered by sponges (mainly Scopalina ruetzleri) and providing refuge to juvenile barracuda; f Cat Cay, close-up of root-fouling community of sponges (orange Scopalina, yellow Mycale) and tunicates (dark green Diplosoma). Plate 1. Marine habitats at the Pelican Cays: a, Cat Cay, red mangrove (Rhizophora mangle) anchored on reef composed mainly of lettuce coral A garicia tenuifolia (photo, Chip Clark), b, close-up of a showing lettuce coral associated with zoanthids, mangrove stilt roots covered by sponges, and schooling fishes; c, outer slope of coral ridge, Cat Cay, composed of lettuce coral and staghorn coral (Acropora cervicornis); d, inner slope of coral ridge, Cat Cay, where lettuce coral is densely populated by sponges and crinoid echinoderms; e, red- mangrove prop roots lining pond E, Fisherman’s Cay, covered by sponges (mainly Scopalina ruetzleri) and providing refuge to juvenile barracuda; f, Cat Cay, close-up of root-fouling community of sponges (orange Scopalina, yellow Mycale) and tunicates (dark green Diplosoma). 330 Plate 2. Examples of sponge habitats and species in the Pelican Cays: a, Chondrilla nucula (brown) and Ulosa funicularis (green) covering dead coral on shallow (2 m) outer reef ridge, Cat Cay; 5, large barrel sponges (Xestospongia muta) on deep (10 m) outer reef ridge, Cat Cay; c, sponge cover on red-mangrove stilt roots, Fisherman’s Cay Pond E, including Desmapsamma anchorata (pink) and Mycale laevis (yellow); d, peat overhang at Manatee Cay, Pond C, with Mycale laevis (yellow) and Spongia tubulifera (black, covered by epizoans); e, close-up of orange sponge Scopalina ruetzleri associated with purple tunicate Clavelina picta, Cat Cay; f, close-up of Callyspongia fallax, Cat Cay. Spi Plate 3. Examples of sponge habitats and species in the Pelican Cays: a, Chondrilla nucula (brown) and Ulosa funicularis (green) covering dead coral on shallow (2 m) outer reef ridge, Cat Cay; b, large barrel sponges (Xestospongia muta) on deep (10 m) outer reef ridge, Cat Cay; c, sponge cover on red-mangrove stilt roots, Fisherman’s Cay Pond E, including Desmapsamma anchorata (pink) and Mycale laevis (yel- low); d, peat overhang at Manatee Cay, Pond C, with Mycale laevis (yellow) and Spongia tubulifera (black, covered by epizoans); e, close-up of orange sponge Scopalina ruetzleri associated with purple tunicate Clavelina picta, Cat Cay; f, close-up of Callyspongia fallax, Cat Cay. 332 Plate 3. Close-ups of representative tunicate species from the Pelican Cays (compare Table 1): a, Didemnum conchyliatum, b, Diplosoma listerianum; c, Clavelina picta; d, Clavelina puertosecensis; e, Ecteinascidia turbinata; f, Microcosmus exasperatus. Plate 2. Close-ups of representative tunicate species from the Pelican Cays (compare Table 1): a, Didemnum conchyliatum; b, Diplosoma lisierianum; c, Clavelina picta; d, Clavelina puertosecensis; e, Ecteinascidia turbinata; f, Microcosmus exasperatus. ATOLL RESEARCH BULLETIN NOS. 466-480 NO. 466. NO. NO. NO. NO. NO. NO. NO. NO. NO. NO. NO. NO. NO. NO. 467. 468. 469. 470. 471. 472. 473. 474. 475. 476. 477. 478. 479. 480. ORIGIN OF THE PELICAN CAYS PONDS, BELIZE BY IAN G. MACINTYRE, WILLIAM F. PRECHT, AND RICHARD B. ARONSON A GENERAL BIOLOGICAL AND GEOLOGICAL SURVEY OF THE RIMS OF PONDS IN THE MAJOR MANGROVE ISLANDS OF THE PELICAN CAYS, BELIZE BY IAN G. MACINTYRE, IVAN GOODBODY, KLAUS RUTZLER, DIANE S. LITTLER, AND MARK M. LITTLER MANGROVE PEAT ANALYSIS AND RECONSTRUCTION OF VEGETATION HISTORY AT THE PELICAN CAYS, BELIZE BY KAREN L. MCKEE AND PATRICIA L. FAULKNER PRELIMINARY HYDROGRAPHIC SURVEYS OF SOME PONDS IN THE PELICAN CAYS, BELIZE BY DANIEL W. URISH HYDROGRAPHY OF A SEMI-ENCLOSED MANGROVE LAGOON, MANATEE CAY, BELIZE BY TRACY A. VILLAREAL, STEVE L. MORTON, AND GEORGE B. GARDNER COMMUNITY STRUCTURE, WATER COLUMN NUTRIENTS, AND WATER FLOW IN TWO PELICAN CAYS PONDS, BELIZE BY THOMAS A. SHYKA AND KENNETH P. SEBENS PHYTOPLANKTON ECOLOGY AND DISTRIBUTION AT MANATEE CAY, PELICAN CAYS, BELIZE BY STEVE L. MORTON DINOFLAGELLATE ASSOCIATIONS IN A CORAL REEF-MANGROVE ECOSYSTEM: PELICAN AND ASSOCIATED CAYS, BELIZE BY MARIA A. FAUST CHECKLIST OF MARINE ALGAE AND SEAGRASSES FROM THE PONDS OF THE PELICAN CAYS, BELIZE BY DIANE S. LITTLER, MARK M. LITTLER, AND BARRETT L. BROOKS EPIPHYTIC FORAMINIFERA OF THE PELICAN CAYS, BELIZE: DIVERSITY AND DISTRIBUTION BY SUSAN L. RICHARDSON DIVERSITY OF SPONGE FAUNA IN MANGROVE PONDS, PELICAN CAYS, BELIZE BY KLAUS RUTZLER, MARIA CHRISTINA DIAZ, ROB W.M. VAN SOEST, SVEN ZEA, KATHLEEN P. SMITH, BELINDA ALVAREZ, AND JANIE WULFF SPONGE PREDATORS MAY DETERMINE DIFFERENCES IN SPONGE FAUNA BETWEEN TWO SETS OF MANGROVE CAYS, BELIZE BARRIER REEF BY JANIE L. WULFF GNATHOSTOMULIDA IN THE PELICAN CAYS, BELIZE BY WOLFGANG STERRER ECHINODERMS OF THE RHOMBOIDAL CAYS, BELIZE: BIODIVERSITY, DISTRIBUTION, AND ECOLOGY BY GORDON HENDLER AND DAVID L. PAWSON DIVERSITY AND DISTRIBUTION OF ASCIDIANS (TUNICATA) IN THE PELICAN CAYS, BELIZE BY IVAN GOODBODY vj ISSUED BY NATIONAL MUSEUM OF NATURAL HISTORY SMITHSONIAN INSTITUTION WASHINGTON, D.C., U.S.A. MARCH 2000 ra) co Bo begt ty a2 ¢ 9.329) Heettcpvaes SMITHSONIAN INSTITUTION LI i oS — os Minn 01375 3942 ree at reevcteia peg ceins a fad re Braet sn tetaclitsetes 6 Docrenrnig thor Ssoncintemceianeitlese 9 viet niet nota ace es tenes Drm Pt ht itet aoe arte 1 NRC SeeL A Reeredt fa tnee ae 6 Rana ompene Po Moats iehaiaeeCtangh ivqasts eee 1) % eG i Plea eee % + ms A tad tlamasemeneannn ati ga bs mW Sa tbraphusnac mite stceget SahD ne bivi beng Sete cane jib tat et cool ene caves eld un Piuchetgl et % TRL Verte rag Seana Pasased. Hebi vata ane Neovaastiaher test bebeves Cate nono Sihate spices eo ii ovgt nnaaeae ipod rye : a eres ye ecenee ORC a TEP ATERCNEF AS Ctr Sea MBS H sith tarot Sha a . ate tehaeye Pe Ree Sheaiceiaes else ee st Shae srwatinese) ee a i th ei bp po % aa fess Con ae cmasergey i Pee pane wat U RON Fr reatenns Linnean Snqaeen eet ; East eelevariestes ea nant ps itiad Helin iterate yeane eee Pascal Reviye wet Mes et eee ee ae mtaine cath ee BAe Oe ania uc finde, RARE neh ieee pater} ROLLS Aeon ely Bt Lara hts