Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-vpsfw Total loading time: 0 Render date: 2024-07-21T09:57:00.206Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  24 August 2017

Imants G. Priede
Affiliation:
University of Aberdeen
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Deep-Sea Fishes
Biology, Diversity, Ecology and Fisheries
, pp. 415 - 469
Publisher: Cambridge University Press
Print publication year: 2017

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Aarestrup, K., Økland, F., Hansen, M.M., Righton, D., Gargan, P., Castonguay, M., Bernatchez, L., Howey, P., Sparholt, H., Pedersen, M. I., & McKinley, R.S. (2009). Oceanic spawning migration of the European eel (Anguilla anguilla). Science 325:1660. doi: 10.1126/science.1178120CrossRefGoogle ScholarPubMed
Agnew, D.J., Heaps, L., Jones, C., Watson, A. Berkieta, K., & Pearce, J. (1999). Depth distribution and spawning pattern of Dissostichus eleginoides at South Georgia. CCAMLR Science, 6, 1936.Google Scholar
Aguayo-Hernández, M. (1995). Biology and fisheries of Chilean hakes (M. gayi and M. australis). In Alheit, J. & Pitcher, T.J. (eds.) Hake: Biology, Fisheries and Markets. London: Chapman & Hall, pp. 305337.CrossRefGoogle Scholar
Ainsley, S.M., Ebert, D.A., & Cailliet, G.M. (2011). Age, growth, and maturity of the whitebrow skate, Bathyraja minispinosa, from the eastern Bering Sea. ICES Journal of Marine Science 68: 14261434. doi:10.1093/icesjms/fsr072CrossRefGoogle Scholar
Akhilesh, K.V., Ganga, U., Pillai, N.G.K., Vivekanandan, E., Bineesh, K.K. Shanis, C.P.R. & Hashim, M. (2011). Deep-sea fishing for chondrichthyan resources and sustainability concerns – a case study from the southwest Indian Coast. Indian Journal of Geo-Marine Sciences 40: 347355.Google Scholar
Alcock, A.W. (1899). A Descriptive Catalogue of the Indian Deep-Sea Fishes in the Indian Museum, Being a Revised Account of the Deep-Sea Fishes Collected by the Royal Indian Marine Survey Ship Investigator. Calcutta: Trustees of the Indian Museum.Google Scholar
Aldrovandi, U. (1613). De piscibus libri V, et de cetis liber unus. Bononia: Bellgambam.Google Scholar
Alexander, R.M. (1966). Physical aspects of swimbladder function. Biological Reviews 41: 141176.CrossRefGoogle ScholarPubMed
Alexander, R.M. (1975). The Chordates. Cambridge: Cambridge University Press.Google Scholar
Alexander, R.M. (1990). Size, speed and buoyancy adaptations in aquatic animals. American Zoologist 30: 189196.CrossRefGoogle Scholar
Alexander, R.M. (1999). Energy for Animal Life. Oxford: Oxford University Press.CrossRefGoogle Scholar
Alheit, J. & Pitcher, T.J. (1995). Hake: Biology, Fisheries and Markets. London: Chapman & Hall.CrossRefGoogle Scholar
Ali, H.A., Mok, H.-K. & Fine, M.L. (2016). Development and sexual dimorphism of the sonic system in deep sea neobythitine fishes: The upper continental slope. Deep-Sea Research I 115: 293308.CrossRefGoogle Scholar
Allain, V., Biseau, A. & Kergoat, B. (2003). Preliminary estimates of French deepwater fishery discards in the Northeast Atlantic Ocean. Fisheries Research, 60, 185–92.CrossRefGoogle Scholar
Allen, G. H. (2001). The Ragfish, Icosteus aenigmaticus Lockington, 1880: A synthesis of historical and recent records from the North Pacific Ocean and the Bering Sea. Marine Fisheries Review 63(4): 131.Google Scholar
Amaoka, K. (2005). Ateleopodiformes: Ateleopodidae: Jellynoses. In: Richards, W.J. (ed.) Early Stages of Atlantic Fishes: An Identification Guide for the Western Central North Atlantic, pp. 295300. Volume 1 Boca Ratan, FL: CRC Press, 2672 pages.Google Scholar
Amaoka, K. & Parin, N.V. (1990). A new flounder, Chascanopsetta megagnatha, from the Sala-y-Gomez Submarine Ridge, Eastern Pacific Ocean (Teleostei: Pleuronectiformes: Bothidae). Copeia 1990(3): 717722.Google Scholar
Amaoka, K. & Yamamoto, E. (1984). Review of the genus Chascanopsetta, with the description of a new species. Bulletin of the Faculty of Fisheries. Hokkaido University 35(4): 201224Google Scholar
Ambrose, D.A. (1996a). Alepocephalidae: Slickheads. In: Moser, H.G. (ed.) The Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 224233. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Ambrose, D.A. (1996b). Anotopteridae: Daggertooth. In: Moser, H.G. (ed.) The Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 369371. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Ambrose, D.A. (1996c). Bythitidae: Brotulas. In: Moser, H.G. (ed.) The Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 538545. Atlas No. 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Ambrose, D.A. (1996d). Carapidae: Pearlfishes. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 532537. Atlas No 33.La Jolla, CA: South West Fisheries Centre.Google Scholar
Ambrose, D.A. (1996e). Macrouridae: Grenadiers. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 483499. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Amemiya, C.T., Alföldi, J., Lee, A.P., Fan, S., Philippe, H. MacCallum, I., Braasch, I., Manousaki, T., Schneider, I., et al. (2013). The African coelacanth genome provides insights into tetrapod evolution. Nature 496(7445): 311316.Google Scholar
Amemiya, C.T., Dorrington, R., & Meyer, A. (2014). The coelacanth and its genome (Introduction to Special Issue: Genome of the African Coelocanth). Journal of Experimental Zoology Part B Molecular and Developmental Evolution 322B: 317321.CrossRefGoogle Scholar
Anastasopoulou, A. & Kapiris, K. (2008). Feeding ecology of the shortnose greeneye Chlorophthalmus agassizi Bonaparte, 1840 (Pisces: Chlorophthalmidae) in the eastern Ionian Sea (eastern Mediterranean). Journal of Applied Ichthyology 24: 170179Google Scholar
Anastasopoulou, A., Yiannopoulos, C., Megalofonou, P., & Papaconstantinou, C. (2006). Distribution and population structure of the Chlorophthalmus agassizi (Bonaparte, 1840) on an unexploited fishing ground in the Greek Ionian Sea. Journal of Applied Ichthyology 22: 521529.Google Scholar
Anderson, M.E. (1986). Family No 95: Parabrotulidae. In: Smith, M.M. & Heemstra, P.C. (eds.) Smith’s Sea Fishes. Berlin: Springer Verlag.Google Scholar
Anderson, M.E. (1989). Review of the eelpout genus Pachycara Zugmayer, 1911 (Teleostei: Zoarcidae), with description of six new species. Proceedings of the California Academy of Sciences 46: 221242.Google Scholar
Anderson, M.E. (2006). Studies on the Zoarcidae (Teleostei: Perciformes) of the Southern Hemisphere. XI. A new species of Pyrolycus from the Kermadec Ridge. Journal of the Royal Society of New Zealand 36(2): 6368.Google Scholar
Anderson, M.E., Crabtree, R.E., Carter, H.J., Sulak, K.J., & Richardson, M.D. (1985). Distribution of demersal fishes of the Caribbean Sea found below 2,000 meters. Bulletin of Marine Science 37: 794807.Google Scholar
Anderson, M.E. & Federov, V.V. (2004). Family Zoarcidae, Swainson 1839 eelpouts. California Academy of Sciences, Annotated Checklists of Fishes 34: 158.Google Scholar
Anderson, M.E. & Peden, A.E. (1988). The eelpout genus Pachycara (Teleostei: Zoarcidae) in the north-eastern Pacific Ocean, with descriptions of two new species. Proceedings of the California Academy of Sciences 46: 8394.Google Scholar
Anderson, M.E., Somerville, R., & Copley, J.T. (2016). A new species of Pachycara Zugmayer, 1911 (Teleostei: Zoarcidae) from deep-sea chemosynthetic environments in the Caribbean Sea. Zootaxa 4066(1): 071077.Google Scholar
Anderson, T.R. & Rice, T. (2006). Deserts on the sea floor: Edward Forbes and his azoic hypothesis for a lifeless deep ocean. Endeavour 30: 131137.CrossRefGoogle Scholar
Anderson, W.D. & Springer, V.G. (2005). Review of the perciform fish genus Symphysanodon Bleeker (Symphysanodontidae), with description of three new species, S. mona, S. parini, and S. rhax. Zootaxa 996: 144.Google Scholar
Andriashev, A.P. (1953). Ancient Deep-Water and Secondary Deep-Water Fishes and Their Importance in Zoogeographical Analysis. Notes of Special Problems in Ichthyology. Moscow, Leningrad: Akademie. Nauk. SSSR. Ikhiol.Kom. [Translation by A. R. Gosline, Ichthyological Laboratory, Bureau of Commercial Fisheries. Fish & Wildlife Service Translation Series 6. Washington DC: US National Museum].Google Scholar
Andriashev, A.P. (1955). On a new Liparid fish from a depth over 7000 meters. Trudy Instituta Okeanologii, Akademiya Nauk, SSSR (Proceedings of the Institute of Oceanology, Academy of Sciences, USSR) 12: 340–344 (in Russian).Google Scholar
Andriashev, A.P., (1998). A review of recent studies of Southern Ocean Liparidae (Teleostei: Scorpaeniformes). Cybium 22(3): 255266.Google Scholar
Armstrong, J.D., Bagley, P.M., & Priede, I.G. (1992). Photographic and acoustic tracking observations of the behaviour of the grenadier, Coryphaenoides (Nematonurus) armatus,the eel, Synaphobranchus bathybius, and other abyssal demersal fish in the North Atlantic Ocean. Marine Biology 112: 535544.CrossRefGoogle Scholar
Armstrong, J.D., Johnstone, A.D.F., & Lucas, M.C. (1992). Retention of intragastric transmitters after voluntary ingestion by captive cod, Gadus morhua L. Journal of Fish Biology 40: 135137CrossRefGoogle Scholar
Armstrong, M.J. & Prosch, R.M. (1991). Abundance and distribution of the mesopelagic fish Maurolicus muelleri in the southern Benguela system, South African Journal of Marine Science, 10(1): 1328, doi: 10.2989/02577619109504615CrossRefGoogle Scholar
Artedi, P. (1738). Ichthyologia, Sive, Opera Omnia de Piscibus Scilicet Bibliotheca Ichthyologica, Philosophia Ichthyologica, Genera Piscium, Synonymia Specierum, Descriptiones Specierum: Omnia in hoc Genere Perfectiora, Quam Antea Ulla / Posthuma Vindicavit, Recognovit, Coaptavit & Edidit Carolus Linnaeus. Leiden: Wishoff.Google Scholar
Aschliman, N.C., Nishida, M., Miya, M., Inoue, J.G., Rosana, K.M., & Naylor, G.J.P. (2012). Body plan convergence in the evolution of skates and rays (Chondrichthyes: Batoidea). Molecular Phylogenetics and Evolution 63 (2012) 2842.Google Scholar
Ast, J.C. & Dunlap, P.V. (2005). Phylogenetic resolution and habitat specificity of members of the Photobacterium phosphoreum species group. Environmental Microbiology 7(10): 16411654.Google Scholar
Atkinson, D.B. (1995). The biology and fishery of Roundnose grenadier (Coryphaenoides rupestris, Gunnerus, 1765) in the North West Atlantic. In: Hopper, A.G. (ed.) Deep-water Fisheries of the North Atlantic Slope. Dordrecht, Netherlands: Kluwer.Google Scholar
Auster, P. J., Gjerde, K., Heupel, E., Watling, L., Grehan, A., & Rogers, A.D. (2011). Definition and detection of vulnerable marine ecosystems on the high seas: problems with the ‘move-on’ rule. ICES Journal of Marine Science, 68: 254264.Google Scholar
Backus, R.H. (1966). The ‘Pinger’ as an aid in deep trawling. Journal du Conseil Permanent International pour l’Exploration de la Mer 30(2): 270277.CrossRefGoogle Scholar
Backus, R.H., Craddock, J.E., Haedrich, R.L., Shores, D.L., Teal, J.M., Wing, A.S., Mead, G.W., & Clarke, W.D. (1968). Ceratoscopelus maderensis: Peculiar sound-scattering layer identified with this Myctophid fish. Science 160: 991993.Google Scholar
Badcock, J., (1981). The significance of meristic variation in Benthosema glaciale (Pisces, Myctophoidei) and of the species distribution off northwest Africa. Deep-Sea Research, 28A: 14771491.Google Scholar
Badcock, J. & Merrett, N.R. (1976). Midwater fishes in the eastern North Atlantic–I. Vertical distribution and associated biology in 30°N, 23°W, with developmental notes on certain myctophids. Progress in Oceanography 7: 358.CrossRefGoogle Scholar
Badcock, J.R. & Merrett, N.R. (1972). On Argyripnus atlanticus, Maul 1952 (Pisces, Stomiatoidei), with a description of post larval forms. Journal of Fish Biology, 4: 277 287.Google Scholar
Badenhorst, A. (1988). Aspects of the South African longline fishery for kingklip Genypterus capensis and the Cape hakes Merluccius capensis and M. paradoxus. South African Journal of Marine Science 6: 3342. doi: 10.2989/025776188784480708Google Scholar
Bagley, P.M., Smith, A. & Priede, I.G. (1994). Tracking movements of deep demersal fishes, Coryphaenoides (Nematonurus) armatus, Antimora rostrata and Centroscymnus coelolepis in the Porcupine Seabight, N.E. Atlantic Ocean. Journal of the Marine Biological Association of the UK 74: 473480.Google Scholar
Bailey, D.M., Bagley, P.M., Jamieson, A.J., Collins, A.M. & Priede, I.G. (2003). In situ investigation of burst swimming and muscle performance in the deep sea fish Antimora rostrata (Günther, 1878). Journal of Experimental Marine Biology & Ecology 285–6: 295311.CrossRefGoogle Scholar
Bailey, D.M., Collins, M.A., Gordon, J.D.M., Zuur, A.F. & Priede, I.G. (2009). Long-term changes in deep-water fish populations in the North East Atlantic: deeper-reaching effect of fisheries? Proceedings of the Royal Society of London B. 275: 19651969.Google Scholar
Bailey, D.M., Genard, B., Collins, M.A., Rees, F., Unsworth, S.K., Battle, E.J.V., Bagley, P.M., Jamieson, A.J., & Priede, I.G. (2005). High swimming and metabolic activity in the deep-sea eel Synaphobranchus kaupii revealed by integrated in situ and in vitro measurements. Physiological & Biochemical Zoology 78: 335346.CrossRefGoogle ScholarPubMed
Bailey, D.M. Jamieson, A.J., Bagley, P.M., Collins, M.A., & Priede, I.G. (2002). Measurement of in situ oxygen consumption of deep-sea fish using an autonomous lander vehicle. Deep Sea Research Part I: Oceanographic Research 49(8): 15191529.Google Scholar
Bailey, D.M., King, N.J., & Priede, I.G. (2007). Cameras and carcasses: Historical and current methods for using artificial food falls to study deep-water animals. Marine Ecology Progress Series, 350: 179191, doi: 10.3354/meps07187CrossRefGoogle Scholar
Bailey, D.M., Ruhl, H.A., & Smith, K.L. Jr. (2006). Long-term change in benthopelagic fish abundance in the abyssal Northeast Pacific Ocean. Ecology. 87(3): 549555.Google Scholar
Bailey, D.M., Wagner, H.-J., Jamieson, A.J., Ross, M.F., & Priede, I.G. (2007). A taste of the deep-sea: The roles of gustatory and tactile searching behaviour in the grenadier fish Coryphaenoides armatus. Deep-sea Research I 54; 99108.CrossRefGoogle Scholar
Baird, R.C. & Hopkins, T.L. (1981). Trophodynamics of the Fish Valenciennellus tripunctulatus. II. Selectivity, Grazing Rates and Resource Utilization. Marine Ecology Progress Series 5: 1119.Google Scholar
Baird, R.C. & Jumper, G.Y. (1993). Olfactory organs in the deep sea hatchetfish Sternoptyx diaphana (Stomiiformes, Sternoptychidae) Bulletin of Marine Science 53(3): 11631167.Google Scholar
Baird, R.C., Jumper, G.Y., & Gallaher, E.E. (1990). Sexual dimorphism and demography in two species of oceanic midwater fishes (Stomiiformes: Sternoptychidae) from the eastern Gulf of Mexico. Bulletin of Marine Science 47: 561566.Google Scholar
Baird, R.C., Wilson, D.F., & Milliken, D.M. (1973). Observations on Bregmaceros nectabanus Whitley in the anoxic, sulfurous water of the Cariaco Trench. Deep-Sea Research 20: 503504.Google Scholar
Baker, K.D., Haedrich, R.L., Snelgrove, P.V.R., Wareham, V.E., Edinger, E.N., & Gilkinson, K.D. (2012b). Small-scale patterns of deep-sea fish distributions and assemblages of the Grand Banks, Newfoundland continental slope. Deep-Sea Research I 65: 171188.Google Scholar
Baker, K.D., Haedrich, R.L., Fifield, D.A., & Gilkinson, K.D. (2012). Grenadier abundance examined at varying spatial scales in deep waters off Newfoundland, Canada, with special focus on the influence of corals. Journal of Ichthyology 52: 678689.CrossRefGoogle Scholar
Baker, K.D., Haedrich, R.L., Snelgrove, P.V.R., Wareham, V.E., Edinger, E.N., & Gilkinson, K.D. (2012a). Small-scale patterns of deep-sea fish distributions and assemblages of the Grand Banks, Newfoundland continental slope. Deep-Sea Research II 65: 171188.Google Scholar
Baker, L.L., Wiff, R., Quiroz, J.C., Flores, A., Céspedes, R., Barrientos, M.A., Ojeda, V., & Gatica, C. (2014). Reproductive ecology of the female pink cusk-eel (Genypterus blacodes): evaluating differences between fishery management zones in the Chilean austral zone. Environmental Biology of Fishes 97: 10831093. doi: 10.1007/s10641-013–0199-2CrossRefGoogle Scholar
Baldwin, C.C. & Johnson, G.D. (1996). Interrelationships of Aulopiformes. In: Stiassny, M.L.J., Parenti, L.R., & Johnson, G.D. (eds.) Interrelationships of Fishes, pp. 355404. San Diego, CA: Academic Press.CrossRefGoogle Scholar
Ballara, S.L. (2014). Fishery characterisation and standardised CPUE analyses for lookdown dory, Cyttus traversi (Hutton, 1872) (Zeidae), 1989–90 to 2011–12. New Zealand Fisheries Assessment Report 2014/62.Google Scholar
Balushkin, A.V. (1996). A new genus and species of liparid fish Palmoliparis beckeri from the northern Kurile Islands (Scorpaeniformes, Liparidae) with consideration of phylogeny of the family. Journal of Ichthyology. 36: 281287.Google Scholar
Balushkin, A.V. (2012). Volodichthys gen. nov. New species of the primitive snailfish (Liparidae: Scorpaeniformes) of the Southern Hemishpere. Description of vew species V. solovjevae sp. nov. (Cooperation Sea, the Antarctic). Journal of Ichthyology 52(1): 110.CrossRefGoogle Scholar
Balushkin, A.V. & Prirodina, V.P. (2010). A new species of Muraenolepididae (Gadiformes) Muraenolepis evseenkoi sp. nova from continental seas of Antarctica. Journal of Ichthyology 50(7): 495502.Google Scholar
Balushkin, A.V. & Voskoboinikova, O.S. (2008). Revision of the genus Genioliparis Andriashev et Neelov (Liparidae, Scorpaeniformes) with description of a new species G. kafanovi sp. n. from the Ross Sea (Antractica). Journal of Ichthyology 48(3): 201208.CrossRefGoogle Scholar
Bambach, RK (2006). Phanerozoic biodiversity mass extinctions. Annual Review of Earth and Planetary Sciences 34: 127155.Google Scholar
Bambach, R.K., Knoll, A.H., & Wang, S.C. (2004). Origination, extinction, and mass depletions of marine diversity. Paleobiology 30(4): 522542.2.0.CO;2>CrossRefGoogle Scholar
Baranes, A. & McCormack, C. (2009). Iago omanensis. The IUCN Red List of Threatened Species 2009: e.T161501A5437914. http://dx.doi.org/10.2305/IUCN.UK.2009-2.RLTS.T161501A5437914.en. Downloaded on 21 March 2017.CrossRefGoogle Scholar
Barange, M., Pakhomov, E.A., Perissinotto, R., Froneman, P.W., Verheye, H.M., Taunton-Clark, J., & Lucas, M.I. (1998). Pelagic community structure of the subtropical convergence region south of Africa and in the mid-Atlantic ocean. Deep-Sea Research I, 45: 16631687.CrossRefGoogle Scholar
Bardack, D. (1991). First fossil hagfish (Myxinoidea): a record from the Pennsylvanian of Illinois. Science 254: 701703.CrossRefGoogle ScholarPubMed
Baremore, I.E. (2010). Reproductive aspects of the Atlantic angel shark Squatina dumeril. Journal of Fish Biology 76: 1682–95. doi: 10.1111/j.1095–8649.2010.02608.x.CrossRefGoogle ScholarPubMed
Barham, A.G., Ayer, N.J., & Boyce, R.E. (1967). Macrobenthos of the San Diego Trough: Photographic census and observations from bathyscaphe, Trieste. Deep-Sea Research 14: 773784.Google Scholar
Barham, E.G. (1966). Deep scattering layer migration and composition: Observations from a Diving Saucer Science, 151: 13991403.Google Scholar
Barnett, M.A. (1984). Mesopelagic fish zoogeography in the central tropical and subtropical Pacific Ocean: Species composition and structure at representative locations in three ecosystems Marine Biology 82: 199208. doi: 10.1007/BF00394103CrossRefGoogle Scholar
Barreiros, J.P., Machado, L.F., Vieira, R.P., & Porteiro, F.M. (2011). Occurrence of Grammicolepis brachiusculus Poey, 1873 (Pisces: Grammicolepididae) in the Azores Archipelago. Life and Marine Sciences 28: 8388.Google Scholar
Bartlett, J.M.S. & Stirling, D. (2003). A short history of the polymerase chain reaction. PCR Protocols  3–6. doi:10.1385/1–59259-384–4:3.Google Scholar
Bassot, J.-M. (1966). On the comparative morphology of some luminous organs. In: Johnson, F.H. &.Haneda, Y. (eds.) Bioluminescence in Progress, pp. 557610. Princeton NJ: Princeton University Press.Google Scholar
Batista, I., Pires, C., Bandarra, N.M., & Goncalves, A. (2001). Chemical characterization and preparation of salted minces from bigeye grunt and longfin bonefish. Journal of Food Biochemistry 25: 527540.CrossRefGoogle Scholar
Beamish, R.J., Leask, K.D., Ivanov, O.A., Balanov, A.A., Orlov, A.M., & Sinclair, B. (1999). The ecology, distribution, and abundance of midwater fishes of the Subarctic Pacific gyres. Progress in Oceanography 43: 399442.CrossRefGoogle Scholar
Beaulieu, S.E. (2002). Accumulation and fate of phytodetritus on the sea floor. Oceanography and Marine Biology Annual Review. 40: 171232.Google Scholar
Becker, J.J., Sandwell, D.T., Smith, W.H.F., Braud, J., Binder, B., Depner, J., Fabre, D., Factor, J., Ingalls, S., Kim, S.-H., Ladner, R., Marks, K., Nelson, S., Pharaoh, A., Trimmer, R., von Rosenberg, J., Wallace, G., & Weatherall, P., (2009). Global bathymetry and elevation data at 30 arc-seconds resolution: SRTM30 PLUS. Marine Geodesy 32: 355371.CrossRefGoogle Scholar
Becker, T.W & Faccenna, C. (2009). A review of the role of subduction dynamics for regional and global plate motions. In: Lallemand, S. & Funiciello, F. (eds.) Subduction Zone Geodynamics, Berlin Heidelberg: Springer-Verlag.Google Scholar
Beebe, W. (1932). Nineteen new species and four post-larval deep-sea fish. Zoologica 13(4): 47107.Google Scholar
Beebe, W. (1934). Half Mile Down. New York: Harcourt, Brace & Co.Google Scholar
Béguer-Pon, M., Castonguay, M., Shan, S., Benchetrit, J,. & Dodson, J.J. (2015). Direct observations of American eels migrating across the continental shelf to the Sargasso Sea. Nature Communications 6:8705 doi: 10.1038/ncomms9705CrossRefGoogle Scholar
Ben-Avraham, Z., Woodside, J., Lodolo, E., Gardosh, M., Grasso, M., Camerlenghi, A., & Vai, G.B. (2006). Eastern Mediterranean basin systems. Geological Society, London, Memoirs 32: 263276. doi: 10.1144/GSL.MEM.2006.032.01.15CrossRefGoogle Scholar
Benfield, M.C., Caruso, J.H., & Sulak, K.J. (2009). In situ video observations of two manefishes (Perciformes: Caristiidae) in the mesopelagic zone of the Northern Gulf of Mexico. Copeia 2009: 637641.Google Scholar
Benn, A.R., Weaver, P.P., Billet, D.S.M., van den Hove, S., Murdock, A.P., Doneghan, G.B., & Le Bas, T. (2010). Human activities on the deep seafloor in the North East Atlantic: An assessment of spatial extent. PLoS ONE 5(9): e12730. doi:10.1371/journal.pone.0012730CrossRefGoogle ScholarPubMed
Benton, M. (2005). Vertebrate palaeontology, 3rd edition. Oxford: Blackwell.Google Scholar
Berelson, W. (2002). Particle settling rates increase with depth in the ocean. Deep-Sea Research II 49: 237251.CrossRefGoogle Scholar
Berenbrink, M., Koldkjær, P., Kepp, O., & Cossins, A.R. (2005). Evolution of oxygen secretion in fishes and the emergence of a complex physiological system. Science 307: 17521757.CrossRefGoogle ScholarPubMed
Bergstad, O.A. (1995). Age determination of deep-water fishes: experiences, status and challenges for the future. In: Hopper, A.G., ed. Deep-water Fisheries of the North Atlantic Oceanic Slope, pp. 267283. NATO ASI Series, Series E: Applied Sciences, Vol. 296. London: Kluwer Academic Publishers.CrossRefGoogle Scholar
Bergstad, O.A., Bjelland, O., & Gordon, J.D.M. (1999). Fish communities on the slope of the eastern Norwegian Sea. Sarsia 84: 6778.Google Scholar
Bergstad, O.A., Clark, L., Hansen, H.Ø., & Cousins, N. (2012). Distribution, population biology, and trophic ecology of the deepwater demersal fish Halosauropsis macrochir (Pisces: Halosauridae) on the Mid-Atlantic Ridge. PLoS ONE 7(2): e31493. doi:10.1371/journal.pone.0031493Google Scholar
Bergstad, O.A., Hansen, H.Ø., & Jørgensen, T. (2013). Intermittent recruitment and exploitation pulse underlying temporal variability in a demersal deep-water fish population. ICES Journal of Marine Science. doi: 10.1093/icesjms/fst202CrossRefGoogle Scholar
Bergstad, O.A., Menezes, G., & Høines, Å.S. (2008). Demersal fish on a mid-ocean ridge: Distribution patterns and structuring factors. Deep-Sea Research II 55: 185202.Google Scholar
Bergstad, O.A., Menezes, G.M.M., Høines, Å., Gordon, J.D.M., & Galbraith, J.K. (2012). Patterns of distribution of deepwater demersal fishes of the North Atlantic mid-ocean ridge, continental slopes, islands and seamounts. Deep-Sea Research I 61: 7483.Google Scholar
Bernardino, A.F., Levin, L.A., Thurber, A.R., & Smith, C.R. (2012). Comparative composition, diversity and trophic ecology of sediment macrofauna at vents, seeps and organic falls. PLoS ONE 7(4): e33515. doi:10.1371/journal.one.0033515CrossRefGoogle ScholarPubMed
Bertelsen, E. (1951). The ceratioid fishes. Ontogeny, taxonomy, distribution and biology. Dana Reports 39: 1276.Google Scholar
Bertelsen, E. (1958) A new type of light organ in the deep-sea fish Opisthoproctus. Nature 181: 862863. doi:10.1038/181862b0CrossRefGoogle Scholar
Bertelsen, E. (1982). First records of metamorphosed males of the families Diceratiidae and Centrophrynidae (Pisces: Ceratioidei). Steenstrupia 8(16): 309315.Google Scholar
Bertelsen, E. & Pietsch, T.W. (1998). Revision of the deepsea anglerfish genus Rhynchactis Regan (Lophiiformes: Gigantactinidae), with descriptions of two new species. Copeia 1998(3): 583590.Google Scholar
Bertelsen, E. & Struhsaker, P.J. (1977). The ceratioid fishes of the genus Thaumatichthys: Osteology, relationships and biology. Galathea Reports 14: 740.Google Scholar
Best, A.C.G. & Bone, Q. (1976). On the integument and photophores of the Alepocephalid fishes Xenodermichthys and Photostylus. Journal of the Marine Biological Association of the U.K. 56: 227236.CrossRefGoogle Scholar
Best, P.B. & Photopoulou, T. (2016). Identifying the ‘demon whale-biter’: Patterns of scarring on large whales attributed to a cookie-cutter shark Isistius sp. PLoS ONE 11(4): e0152643. doi:10.1371/journal.pone.0152643CrossRefGoogle ScholarPubMed
Betancur, R R., Broughton, R.E., Wiley, E.O., Carpenter, K., López, J.A., Li, C., Holcroft, N.I., Arcila, D., Sanciangco, M., Cureton, J.C. II, Zhang, F., Buser, T., Campbell, M.A., Ballesteros, J.A., Roa-Varon, A., Willis, S., Borden, W.C., Rowley, T., Reneau, P.C., Hough, D.J., Lu, G., Grande, T., Arratia, G. & Ortí, G. (2013). The tree of life and a new classification of bony fishes. PLOS Currents Tree of Life. 2013 Apr 18. edition 1. doi: 10.1371/currents.tol.53ba26640df0ccaee75bb165c8c26288.CrossRefGoogle Scholar
Beverton, R.J.H. & Holt, S.J. (1957). On the dynamics of exploited fish populations. Fishery Investigation Series II, Vol. XIX.Google Scholar
Bianchi, G., Carpenter, K.E., Roux, J.-P., Molloy, F.J., Boyer, D., & Boyer, H.J. (1999). FAO species identification field guide for fishery purposes. The living marine resources of Namibia. Rome: FAO. http://www.fao.org/docrep/009/x3478e/x3478e00.HTMGoogle Scholar
Bilecenoglu, M., Kaya, M., & Irmak, E. (2006). First records of the slender snipe eel, Nemichthys scolopaceus (Nemichthyidae), and the robust cusk-eel, Benthocometes robustus (Ophidiidae), from the Aegean Sea. Acta Ichthyologica et Piscatoria 36 (1): 8588.Google Scholar
Billett, D.S.M., Bett, B.J., Jacobs, C.L., Rouse, I.P. & Wigham, B.D. (2006). Mass deposition of jellyfish in the deep Arabian Sea. Limnology & Oceanography 51: 20772083.Google Scholar
Billett, D.S.M., Lampitt, R.S., Rice, A.L., & Mantoura, R.F.C. (1983). Seasonal sedimentation of phytoplankton to the deep-sea benthos. Nature 302: 520522 doi:10.1038/302520a0CrossRefGoogle Scholar
Biscoito, M. & Almeida, A.J. (2004). New Species of Pachycara Zugmayer (Pisces: Zoarcidae) from the Rainbow Hydrothermal Vent Field (Mid-Atlantic Ridge). Copeia, 2004 (3): 562568.CrossRefGoogle Scholar
Biscoito, M., Segonzac, M., Almeida, A.J., Desbruyères, D., Geistdoerfer, P., Turnipseed, M., & Van Dover, C. (2002). Fishes from the hydrothermal vents and cold seeps – an update. Cahiers de Biologie Marine 43: 359362.Google Scholar
Blakey, R. (2011). Library of Paleogeography. Colorado Plateau Geosystems, Flagstaff, AZ: Author. http://cpgeosystems.com/paleomaps.html Accessed 20 June 2014.Google Scholar
Blanchard, F. (2001). The effect of fishing on demersal fish community dynamics: an hypothesis. ICES Journal of Marine Science 58: 711718.CrossRefGoogle Scholar
Blankenship, L. E. & Levin, L. A. (2007). Extreme food webs: foraging strategies and diets of scavenging amphipods from the ocean’s deepest 5 kilometers. Limnology and Oceanography 52, 16851697.CrossRefGoogle Scholar
Blaxter, J.H.S. (1979). The effect of hydrostatic pressure on fishes. In: Ali, M.A. (ed.) Environmental Physiology of Fishes. New York: Plenum Press.Google Scholar
Blaxter, J.H.S., Wardle, C.S. & Roberts, B.L. (1971). Aspects of the circulatory physiology and muscle systems of deep-sea fish. Journal of the Marine Biological Association of the UK, 51: 9911006.Google Scholar
Block, B.A., Booth, D.T., & Carey, F.G. (1992). Direct measurement of swimming speeds and depth of blue marlin. Journal of Experimental Biology 166: 267284.CrossRefGoogle Scholar
Block, B.A., Dewar, H., Blackwell, S.B., Williams, T.D. et al. (2001). Migratory movements, depth preferences, and thermal biology of Atlantic bluefin tuna. Science 293: 13101314.Google Scholar
Boehlert, G.W., Wilson, C.D., & Mizuno, K. (1994). Populations of the Sternotychid fish Maurolicus muelleri on seamounts in the Central North Pacific. Pacific Science. 48: 5769.Google Scholar
Boetius, A. & Wenzhöfer, F. (2013). Seafloor oxygen consumption fuelled by methane from cold seeps. Nature Geoscience 6: 725734.Google Scholar
Bone, Q. (1972). Buoyancy and hydrodynamic functions of integument in the castor oil fish, Ruvettus pretiosus (Pisces: Gempylidae). Copeia 1972: 7887.Google Scholar
Bone, Q. (1973). A note on the buoyancy of some lantern-fishes. (Myctophoidei). Journal of the Marine Biological Association. U.K. 53:619633Google Scholar
Bone, Q. & Moore, R.H. (2008). Biology of Fishes, 3rd edition. Abingdon. England: Taylor & Francis.Google Scholar
Bone, Q. & Roberts, BL. (1969). The density of elasmobranchs. Journal of the Marine Biological Association U K. 49: 913937. doi: 10.1017/s0025315400038017Google Scholar
Bourret, P. (1985). Poissons Téléostéens: Gonostomatidae, Sternoptychidae, et Myctophidae (MUSORSTOM II). Mémoires du Muséum National d’Histoire Naturelle., séries. A, Zoologie. Paris. 133: 5582.Google Scholar
Bowering, W.R. & Brodie, W.B. (1995). Greenland halibut (Reinhardtius hippoglossoides). A review of the dynamics of its distribution and fisheries off Eastern Canada and Greenland. In: Hopper, A.G. (ed.) Deep-Water Fisheries of the North Atlantic Slope, pp. 113160. Dordrecht, Netherlands: Kluwer.CrossRefGoogle Scholar
Braccini, J.M. (2008). Feeding ecology of two high-order predators from south-eastern Australia: The coastal broadnose and the deepwater sharpnose sevengill sharks. Marine Ecology Progress Series 371: 273284. doi:10.3354/meps07684.Google Scholar
Braccini, J.M., Gillanders, B.M. & Walker, T.I. (2006). Hierarchical approach to the assessment of fishing effects on non-target chondrichthyans: case study of Squalus megalops in southeastern Australia. Canadian Journal of Fisheries and Aquatic Sciences 63 (11): 24562466 doi: 10.1139/f06-141CrossRefGoogle Scholar
Bradbury, M.G. (1988). Rare fishes of the deep-sea genus Halieutopsis: a review with descriptions of four new species (Lophiiformes: Ogcocephalidae). Fieldiana 44: 122Google Scholar
Bradbury, M.G. (1999). A review of the fish genus Dibranchus with descriptions of new species and a new genus, Solocisquama (Lophiiformes, Ogcocephalidae). Proceedings of the California Academy of Sciences 51(5): 259310.Google Scholar
Brander, K. (1981). Disappearance of common skate Raia batis from Irish Sea. Nature 290: 4849.CrossRefGoogle Scholar
Brauer, A. (1906). Die Tiefsee-Fische. I. Systematischer Teil. Wissenschaftliche Ergebnisse der deutschen Tiefsee-Expedition ‘Valdivia’, 1898–99 (editor C. Chun) Volume. 15(1): 1–432, Gustav Fischer, Jena.Google Scholar
Brauer, A. (1908). Die Tiefsee-Fische. II. Anatomischer Teil. Wissenschaftliche Ergebnisse der deutsch fsee-Expedition ‘Valdivia’, 1898–99 (editor C. Chun) Volume. 15(2): 1–266, Gustav Fischer, Jena.Google Scholar
Bray, D.J. (2011). Toothed Whiptail, Lepidorhynchus denticulatus, in Fishes of Australia, accessed 27 Jul 2015, www.fishesofaustralia.net.au/home/species/4355Google Scholar
Briggs, J.C (2003). Marine centres of origin as evolutionary engines. Journal of Biogeography 30: 118.CrossRefGoogle Scholar
Bray, R.A. (2004). The bathymetric distribution of the digenean parasites of deep-sea fishes. Folia Parasitologica 51: 268274.Google Scholar
Bray, D.J. & Gomon, M.F. (2011). Atlantic Sabretooth, Coccorella atlantica, in Fishes of Australia, accessed 22 May 2015, www.fishesofaustralia.net.au/home/species/2796Google Scholar
Briggs, J.C. (1960). Fishes of worldwide (circumtropical) distribution. Copeia, 171–180.CrossRefGoogle Scholar
Britton, J.C. & Morton, B. (1994). Marine carrion and scavengers. Oceanography and Marine Biology: An Annual Review 32: 369434.Google Scholar
Brodeur, R.D., Seki, M.P., Pakhomov, E.A., & Suntsov, A.V. (2005). Micronekton – What are they and why are they important? PICES Press 13(1): 711. Online publication accessed 18 July 2016 www.pices.int/publications/pices_pressGoogle Scholar
Broecker, W.S. (1991). The great ocean conveyor. Oceanography 4: 7989.CrossRefGoogle Scholar
Brown, S.K., Mahon, R., Zwanenburg, K.C.T, Buja, K.R., Claflin, L.W., O’Boyle, R.N., Atkinson, B., Sinclair, M., Howell, G., & Monaco, M.E. (1996). East Coast of North America Groundfish: Initial Explorations of Biogeography and Species Assemblages. Silver Spring, MD: National Oceanic and Atmospheric Administration, and Dartmouth, NS: Department of Fisheries and Oceans.Google Scholar
Bruland, K.W. & Silver, M. W. (1981). Sinking rates of fecal pellets from gelatinous zooplankton (salps, pteropods, doliolids). Marine Biology 63: 295300, doi: 10.1007/BF00395999CrossRefGoogle Scholar
Brulla, L., Nizet, E., & Verney, E.B. (1953). Blood perfusion of the kidney of Lophius Piscatorius L. Journal of the Marine Biological Association of the UK 32: 329336.Google Scholar
Bruun, A.F. (1951). The Philippine Trench and its bottom fauna. Nature 168: 692693 doi: 10.1038/168692b0CrossRefGoogle Scholar
Buckland, S.T., Anderson, D.R., Burnham, K.P., Laake, J.L., Borchers, D.L., & Thomas, L. (2001). Introduction to Distance Sampling: Estimating Abundance of Biological Populations. Oxford: Oxford University Press.CrossRefGoogle Scholar
Buesseler, K.O. Lamborg, C.H., Boyd, P.W., Lam, P.J., Trull, T.W., Bidigare, R.R., Bishop, J.K.B., Casciotti, K.L., Dehairs, F., Elskens, M., Honda, M., Karl, D.M., Siegel, D.A., Silver, M.W., Steinberg, D.K., Valdes, J., Van Mooy, B., & Wilson, S. (2007). Revisiting carbon flux through the ocean’s twilight zone. Science 316: 567570.Google Scholar
Buhl-Mortensen, L., Vanreusel, A., Gooday, A.J., Levin, L.A., Priede, I.G., Buhl-Mortensen, P., Gheerardyn, H., King, N.J., & Raes, M. (2010). Biological structures as a source of habitat heterogeneity and biodiversity on the deep ocean margins. Marine Ecology 31: 2150.CrossRefGoogle Scholar
Bullough, L.W., Turrell, W.R., Buchan, P. & Priede, I.G. (1998). Commercial deep water trawling at sub-zero temperatures – observations from the Faroe-Shetland channel. Fisheries Research 39: 3341.Google Scholar
Burne, R.H. (1926). A contribution to the anatomy of the ductless glands and lymphatic system of the angler fish (Lophius piscatorius). Philosophical Transactions of the Royal Society of London. B 215: 157.Google Scholar
Busby, M.S. & Orr, J.W. (1999). A pelagic basslet Howella sherborni (Family Acropomatidae) off the Aleutian Islands. Alaska Fishery Research Bulletin 6(1): 4953.Google Scholar
Bussing, W.A. & López-S, M.I. (1977). Guentherus altivela Osorio, the first ateleopodid fish reported from the eastern Pacific Ocean. Revista de Biologia Tropical 25(2): 179190.Google Scholar
Butler, J.L. & Ahlstrom, E.H. (1976). Review of the deep-sea fish genus Scopelengys (Neoscopelidae) with a description of a new species, Scopelengys clarkei, from the Central Pacific. Fishery Bulletin 74: 142150.Google Scholar
Butler, M., Bollens, S.M., Burkhalter, B., Madin, L.P., & Horgan, E. (2001). Mesopelagic fishes of the Arabian Sea: Distribution, abundance and diet of Chauliodus pammelas, Chauliodus sloani, Stomias affinis, and Stomias nebulosus. Deep-Sea Research II 48: 13691383.CrossRefGoogle Scholar
Byrkjedal, I., Poulsen, J.Y. & Galbraith, J. (2011). Leptoderma macrophthalmum n.sp., a new species of smooth-head (Otocephala: Alepocephalidae) from the Mid Atlantic Ridge. Zootaxa 2876: 4956.Google Scholar
Caira, J.N., Benz, G.W., Borucinska, J., & Kohler, N.E. (1997). Pugnose eels, Simenchelys parasiticus (Synaphobranchidae) from the heart of a shortfin mako, Isurus oxyrinchus (Lamnidae). Environmental Biology of Fishes 49: 139144.CrossRefGoogle Scholar
Calistry. (2017). Van Der Waals equation calculator. http://calistry.org/calculate/vanDerWaalsCalculator. On line resource accessed 9 March 2017.Google Scholar
Cavin, L. & Martin, M. (1995). Les Actinoptérygiens et la limite Crétacé-Tertiaire. Geobios 28 (supplement 2): 183188.Google Scholar
Campbell, M.A., Chen, W.-J., & López, A. (2013). Are flatfishes (Pleuronectiformes) monophyletic? Molecular Phylogenetics and Evolution 69 (2013): 664673.CrossRefGoogle ScholarPubMed
Campbell, R.A., Haedrich, R.L,. & Munroe, T.A. (1980) Parasitism and ecological relationships among deep-sea benthic fishes. Marine Biology 57: 301313.CrossRefGoogle Scholar
Campillo, A. (1992). Les Pecheries Françaises de Mediterranee, Synthese des Connaissances. Réf. RI DRV 92-019 RH/Sete. Accessed online 2 March 2015: http://archimer.ifremer.fr/doc/1992/rapport-1125.pdfGoogle Scholar
Canals, M., Puig, P., Durrieu de Madron, X., Heussner, S., Palanques, A., & Fabrés, J. (2006). Flushing submarine canyons. Nature 444: 354357.Google Scholar
Carrasson, M. & Matallanas, J. (1998). Feeding habits of Alepocephalus rostratus (Pisces: Alepocephalidae) in the Western Mediterranean Sea. Journal of the Marine Biological Association of the U.K. 78: 12951306.CrossRefGoogle Scholar
Carrete, Vega G. & Wiens, J.J. (2012). Why are there so few fish in the sea? Proceedings of the Royal Society of London. B 279: 23232329. doi:10.1098/rspb.2012.0075Google Scholar
Carter, H.J. & Musick, J.A. (1985). Sexual dimorphism in the deep-sea fish Barathrodemus manatinus (Ophidiidae). Copeia 1985: 6973.CrossRefGoogle Scholar
Caruso, J.H. (1989). Systematics and distribution of Atlantic chaunacid anglerfishes (Pisces: Lophiiformes). Copeia 1989(1): 153165.CrossRefGoogle Scholar
Caruso, J.H., Ho, H.-C., & Pietsch, T.W. (2006). Chaunacops Garman, 1899, a senior objective synonym of Bathychaunax Caruso, 1989 (Lophiiformes: Chaunacoidei: Chaunacidae). Copeia (1): 120–121.Google Scholar
Casey, J.M. & Myers, R.A. (1998). Near extinction of a large, widely distributed fish. Science 31: 690692. doi: 10.1126/ science.281.5377.690.Google Scholar
Castle, P.H.J. (1959). A large leptocephalid (Teleostei, Apodes) from off South Westland, New Zealand. Transactions of the Royal Society of New Zealand 87 (pts 1–2): 179184.Google Scholar
Castle, P.H.J. (1991). First Indo-Pacific Record of the Eel Family Myrocongridae, with the Description of a New Species of Myroconger. Copeia 1991: 148150.Google Scholar
Castonguay, L. & McCleave, J.D. (1987a). Distribution of leptocephali of the oceanic species Derichthys serpentinus and Nessorhamphus ingolfianus (Family Derichthyidae) in the western Sargasso Sea in relation to physical oceanography. Bulletin of Marine Science 41: 807821.Google Scholar
Castonguay, M. & McCleave, J.D. (1987b) Vertical distributions, diel and ontogenetic vertical migrations and net avoidance of leptocephali of Anguilla and other common species in the Sargasso Sea. Journal of Plankton Research 9: 195214.Google Scholar
Catul, V., Gauns, M., & Karuppasamy, P.K. (2011). A review on mesopelagic fishes belonging to family Myctophidae. Reviews in Fish Biology and Fisheries 21: 339354.Google Scholar
Causse, R., Biscoito, M., & Briand, P (2005). First record of the deep-sea eel Ilyophis saldanhai (Synaphobranchidae, Anguilliformes) from the Pacific Ocean. Cybium 29: 413416.Google Scholar
Cavin, L. (2001). Effects of the Cretaceous–Tertiary boundary event on bony fishes. In: Buffetaut, E. & Koeberl, C. (eds.) Geological and Biological Effects of Impact Events, pp. 141158. Berlin: Springer.Google Scholar
CCAMLR. (2015). Toothfish fisheries. Commission for the Conservation of Antarctic Marine Living Resources. www.ccamlr.org/en/fisheries/toothfish-fisheries Accessed 10 December 2015.Google Scholar
CEC. (2006). Deep-sea gillnet fisheries. Report of the Scientific, Technical and Economic Committee for Fisheries. Accessed 24 February 2016. https://stecf.jrc.ec.europa.eu/documents/43805/122924/06–11_Adhoc+06-01+Deep-Sea+Gillnet+Fisheries.pdfGoogle Scholar
Cember, R.P. (1988). On the sources, formation, and circulation of Red Sea deep water. Journal of Geophysical Research 93 (C7): 81758191.Google Scholar
Chanet, B., Guintard, C., Betti, E., Gallut, C., Dettaï, A., & Lecointre, G. (2013). Evidence for a close phylogenetic relationship between the teleost orders Tetraodontiformes and Lophiiformes based on an analysis of soft anatomy. Cybium 37(3): 179198.Google Scholar
Chapman, L., Desurmont, A., Choi, Y., Boblin, P., Sokimi, W. & Beverly, S. (2008). Fish species identification manual for deep-bottom snapper fishermen. Secretariat of the Pacific Community, Noumea, New Caledonia.Google Scholar
Charter, S.R. (1996). Nemichthyidae: Snipe Eels. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 122129. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Chave, E.H. & Mundy, B.C. (1994). Deep-sea benthic fish of the Hawaiian Archipelago, Cross Seamount, and Johnston Atoll. Pacific Science 48: 367409.Google Scholar
Chavez, F.P. & Messié, M. (2009). A comparison of Eastern Boundary Upwelling Ecosystems. Progress in Oceanography 83: 8096.CrossRefGoogle Scholar
Chembian, A.J. (2007). New record of Rhinochimaera atlantica (Chimaeriformes: Rhinochimaeridae) spawning ground in the Gulf of Mannar along the south-east coast of India. Indian Journal of Fisheries 54(4): 345350.Google Scholar
Chernova, N.V. & Stein, D.L. (2002). Ten new species of Psednos (Pisces, Scorpaeniformes: Liparidae) from the Pacific and North Atlantic Oceans. Copeia 2002: 755778.Google Scholar
Chernova, N.V. & Stein, D.L. (2004). A remarkable new species of Psednos (Teleostei: Liparidae) from the western North Atlantic Ocean. Fishery Bulletin 102: 245250.Google Scholar
Childress, J.J., Barnes, A.T., Quetin, L.B., & Robison, B.H. (1978). Thermally protecting cod ends for the recovery of living deep-sea animals. Deep-Sea Research 25: 419422.CrossRefGoogle Scholar
Childress, J.J. & Somero, G. (1979). Depth-related enzymic activities in muscle, brain and heart of deep-living pelagic marine teleosts. Marine Biology 52: 273283.Google Scholar
Childress, J.J., Taylor, S.M., Cailliet, G.M. & Price, M.H. (1980). Patterns of growth, energy utilization and reproduction in some meso- and bathypelagic fishes off Southern California. Marine Biology 61: 2740.Google Scholar
Childress, J.J. & Thuesen, E.V. (1995). Metabolic potentials of deep-sea fishes: A comparative approach. In: Hochachka, P.W. & Mommsen, T.P. (eds.) Biochemistry and Molecular Biology of Fishes, pp. 175196, Vol. 5. Amsterdam: Elsevier.Google Scholar
Christiansen, B., Vieira, R.P., Christiansen, S., Denda, A., Oliveira, F., & Gonçalves, J.M.S. (2015). The fish fauna of Ampère Seamount (NE Atlantic) and the adjacent abyssal plain. Helgoland Marine Research 69: 1323. doi:10.1007/s10152-014–0413-4CrossRefGoogle Scholar
Claes, J.M., Ho, H.C., & Mallefet, J. (2012). The control of luminescence from pygmy shark (Squaliolus aliae) photophores. Journal of Experimental Biology 215: 16911699Google Scholar
Claes, J.M., Nilsson, D.-E., Straube, N., Collin, S.P., & Mallefet, J. (2014). Iso-luminance counterillumination drove bioluminescent shark radiation. Scientific Reports 4: 4328. doi:10.1038/srep04328 (2014).Google Scholar
Claes, J.M., Partridge, J.C., Hart, N.S., Garza-Gisholt, E., Ho, H.-C., et al. (2014) Photon Hunting in the Twilight Zone: Visual Features of Mesopelagic Bioluminescent Sharks. PLoS ONE 9(8): e104213. doi:10.1371/journal.pone.0104213CrossRefGoogle ScholarPubMed
Clark, M. (1995). Experience with management of Orange Roughy (Hoplostethus atlanticus) in New Zealand waters, and the effects of commercial fishing on stocks over the period 1980–1993. In: Hopper, A.G. (ed.) Deep-Water Fisheries of the North Atlantic Slope, pp. 151266. Dordrecht, Netherlands: Kluwer.Google Scholar
Clark, M. (2001). Are deep water fisheries sustainable? the example of orange roughy (Hoplostethus atlanticus) in New Zealand. Fisheries Research 51: 123135.CrossRefGoogle Scholar
Clark, M.R., Althaus, F., Schlacher, T.A., Williams, A., Bowden, D.A., & Rowden, A.A. (2015). The impacts of deep-sea fisheries on benthic communities: a review. ICES Journal of Marine Science. doi: 10.1093/icesjms/fsv123.Google Scholar
Clark, M.R., Althaus, F., Williams, A., Niklitschek, E., Menezes, G.M., Hareide, N.-R., Sutton, P., & O’Donnell, C. (2010). Are deep-sea demersal fish assemblages globally homogenous? Insights from seamounts. Marine Ecology 31(Suppl. 1): 3951.CrossRefGoogle Scholar
Clarke, J., Milligan, T.J., Bailey, D.M., & Neat, F.C. (2015). A scientific basis for regulating deep-sea fishing by depth. Current Biology 25: 24252429.CrossRefGoogle ScholarPubMed
Clarke, M.R. (1969). A new midwater trawl for sampling discrete depth horizons. Journal of the Marine Biological Association UK 49: 945960.CrossRefGoogle Scholar
Clarke, M.R. (2003). Searching for deep sea squids. Berliner paläobiologische Abhandlungen 3: 4959.Google Scholar
Clarke, M.W., Kelly, C.J., Connolly, P.L., & Molloy, J.P. (2003). A life history approach to the assessment and management of deepwater fisheries in the North east Atlantic. Journal of Northwest Atlantic Fisheries Science 31: 401411.Google Scholar
Clarke, T.A. (1982). Feeding habits of stomiatoid fishes from Hawaiian waters. Fishery Bulletin 80: 287304.Google Scholar
Clausen, D.M. (2008). The giant grenadier in Alaska. American Fisheries Society Symposium 63: 413450.Google Scholar
Clement, G., Wells, R., & Gallagher, C.M. (2008). Industry management within the New Zealand quota management system: The Orange Roughy Management Company. In: Townsend, R.E., Shotton, R., & Uchida, H. (eds.) Case Studies in Fisheries Self-governance, pp. 277290. FAO Fisheries Technical Paper 504. Rome: FAO.Google Scholar
Coad, B.W. (1995). Encyclopedia of Canadian fishes. Canadian Museum of Nature and Canadian Sportfishing Productions Inc. Singapore.Google Scholar
Coghlan, A. (2008). Deep-sea fishing. In: Lück, E. (ed.) The Encyclopedia of Tourism and Recreation in Marine Environments, pp. 132133. Wallingford, England: CABI.Google Scholar
Cohen, D.H. & Haedrich, R.L. (1983). The fish fauna of the Galapagos thermal vent region. Deep-Sea Research 30A: 371379.Google Scholar
Cohen, D.M. (1977). Swimming performance of the gadoid fish Antimora rostrata at 2400 meters. Deep-Sea Research 24: 275277.Google Scholar
Cohen, D.M. & Rohr, B.A. (1993). Description of a giant Circumglobal Lamprogrammus species (Pisces: Ophidiidae). Copeia 1993(2): 470475.CrossRefGoogle Scholar
Cohen, D.M., Rosenblatt, R.H., & Moser, H.G. (1990). Biology and description of a bythitid fish from deep-sea thermal vents in the tropical eastern Pacific. Deep-Sea Research 37: 267283.Google Scholar
Collett, R. (1880). Fishes. The Norwegian North-Atlantic Expedition 1876–1878. 3(1): 1166.Google Scholar
Collett, R. (1896). Poissons provenant des campagnes du Yacht “L’Hirondelle” (1885–1888). Résultats des campagnes scientifiques accomplies sur son yacht par Albert I, Prince Souverain de Monaco. Résultats des campagnes scientifiques du Prince de Monaco. Monaco: Imprimerie de Monaco.Google Scholar
Collette, B.B., Curtis, M., Smith-Vaniz, W.F., Pina Amargos, F., Williams, J.T., & Grijalba Bendeck, L. (2015). Ruvettus pretiosus. The IUCN Red List of Threatened Species 2015: e.T190432A16644022. http://dx.doi.org/10.2305/IUCN.UK.2015–4.RLTS.T190432A16644022.en. Downloaded on 09 April 2016.Google Scholar
Collin, S.P., Hoskins, R. V., & Partridge, J. C. (1998). Seven retinal specialisations in the tubular eye of the deepsea pearleye, Scopelarchus michaelsarsi: a case study in visual optimisation. Brain Behaviour and Evolution, 51: 291314.CrossRefGoogle Scholar
Collin, S.P. & Partridge, J.C. (1996). Retinal specialisations in the eyes of deep-sea teleosts. Journal of Fish Biology 49A: 157174.CrossRefGoogle Scholar
Collins, M.A., Bailey, D.M., Ruxton, G.D., & Priede, I.G. (2005). Trends in body size across an environmental gradient: a differential response in scavenging and non-scavenging demersal deep-sea fish. Proceedings of the Royal Society B 272: 20512057.CrossRefGoogle ScholarPubMed
Collins, M.A., Brickle, P., Brown, J. & Belchier, M. (2010). The Patagonian Toothfish: biology, ecology and fishery. Advances in Marine Biology 58: 227300.Google Scholar
Collins, M.A., Priede, I.G., & Bagley, P.M. (1999). In situ comparison of activity of two deep-sea scavenging fish occupying different depths zones. Proceedings of the Royal Society, Series B 266:20112016.Google Scholar
Collins, M.A., Stowasser, G., Fielding, S., Shreeve, R., Xavier, J.C., Venables, H.J., Enderlein, P., Cherel, Y., & Van de Putte, A. (2012). Latitudinal and bathymetric patterns in the distribution and abundance of mesopelagic fish in the Scotia Sea. Deep-Sea Research II 59–60: 189198.CrossRefGoogle Scholar
Collins, M.A., Xavier, J.C., Johnston, N.M., North, A.W., Enderlein, P., Tarling, G.A., Waluda, C.M., Hawker, E.J., & Cunningham, N.J. (2008). Patterns in the distribution of myctophid fish in the northern Scotia Sea ecosystem. Polar Biology 31: 837851. DOI 10.1007/s00300-008-0423-2Google Scholar
Colman, J.A. (1995). Biology and fisheries of New Zealand hake (M.australis). In: Alheit, J. & Pitcher, T.J. (eds.) Hake: Biology, Fisheries and Markets, pp. 365388. London: Chapman & Hall.Google Scholar
Company, J.B., Puig, P., Sardà, F., Palanques, A., Latasa, M., et al. (2008). Climate influence on deep sea populations. PLoS ONE 3(1): e1431. doi:10.1371/journal.pone.0001431Google Scholar
Connelly, D.P. et al. (2012). Hydrothermal vent fields and chemosynthetic biota on the world’s deepest seafloor spreading centre. Nature Communications 3, Article number: 620. doi:10.1038/ncomms1636CrossRefGoogle ScholarPubMed
Cook, S.F. & Compagno, L. J.V. (2005). Hexanchus griseus. The IUCN Red List of Threatened Species 2005: e.T10030A3155348. http://dx.doi.org/10.2305/IUCN.UK.2005.RLTS.T10030A3155348.en. Downloaded on 25 February 2016.Google Scholar
Corbella, C. (2013). Taxonomy and phylogeny of Lyconidae (Teleostei:Zoarcidae) from the Southern Ocean and Magellan Province. Doctoral Thesis. Univeritat Autonoma de Barcelona.Google Scholar
Corliss, J.B. & Ballard, R.D., (1977). Oases of life in the cold abyss. National Geographic Magazine 152: 441453.Google Scholar
Correia, A.T., Faria, R., Alexandrino, P., Antunes, C., Isidro, E.J., & Coimbra, J. (2006). Evidence for genetic differentiation in the European conger eel Conger conger based on mitochondrial DNA analysis. Fisheries Science 72: 2027.Google Scholar
Correia, A.T., Ramos, A.A., Barros, F., Silva, G., Hamer, P., Morais, P., Cunha, R.L., & Castilho, R. (2012). Population structure and connectivity of the European conger eel (Conger conger) across the north-eastern Atlantic and western Mediterranean: Integrating molecular and otolith elemental approaches. Marine Biology 159:15091525. doi 10.1007/s00227-012–1936-3Google Scholar
Cossins, A.R. & MacDonald, A.G. (1986). Homeoviscous adaptation under pressure: III The fatty acid composition of liver mitochondrial phospholipids of deep-sea fish. Biochim. Biophys Acta 860: 325335.CrossRefGoogle Scholar
Costello, M.J., Cheung, A. & De Hauwere, N. (2010). Surface area and seabed area, volume, depth slope and topographic variation for the World’s Sea, Oceans and countries. Environmental Science & Technology 44: 88218828.CrossRefGoogle ScholarPubMed
Costello, M.J., May, R.M., & Stork, N.E. (2013). Can we name Earth’s species before they go extinct? Science 339: 413416. doi: 10.1126/science.1230318CrossRefGoogle ScholarPubMed
Costello, M.J., McCrea, M., Freiwald, A., Lundalv, T., Jonsson, L., Bett, B.J., van Weering, T.C.E., de Haas, H., Roberts, J.M. & Allen, D. (2005). Role of cold-water Lophelia pertusa coral reefs as fish habitat in the NE Atlantic. In: Freiwald, A. & Roberts, J.M. (eds.) Cold-Water Corals and Ecosystems. Berlin: Springer-Verlag.Google Scholar
Cousins, N.J., Linley, T., Jamieson, A.J., Bagley, P.M., Blades, H., Box, T., Chambers, R., Ford, A., Shields, M.A., & Priede, I.G. (2013). Bathyal demersal fishes of Charlie-Gibbs fracture zone region (49°–54°N) of the Mid Atlantic Ridge: II. Baited camera lander observations. Deep-Sea Research II 98: 397406. http://dx.doi.org/10.1016/j.dsr2.2013.08.002CrossRefGoogle Scholar
Cousins, N.J. & Priede, I.G. (2012). Abyssal demersal fish fauna composition in two contrasting productivity regions of the Crozet Plateau, Southern Indian Ocean. Deep-Sea Research I 64: 7177. doi:10.1016/j.dsr.2012.02.003Google Scholar
Cousins, N.J., Shields, M.A., Crockard, D. & Priede, I.G. (2013). Bathyal demersal fishes of Charlie Gibbs fracture zone region (49°–54°N) of the Mid-Atlantic Ridge: I. Results from trawl surveys. Deep-Sea Research II 98: 388396. http://dx.doi.org/10.1016/j.dsr2.2013.08.012Google Scholar
Cowie, G.L. & Levin, L.A. (2009). Benthic biological and biogeochemical patterns and processes across an oxygen minimum zone (Pakistan margin, NE Arabian Sea). Deep-Sea Research II 56: 261270.Google Scholar
Cox, G.K., Sandblom, E., Richards, J.G. & Farrell, A.P. (2011). Anoxic survival of the Pacific hagfish (Eptatretus stoutii). Journal of Comparative Physiology B 181, 361371. 10.1007/s00360-010–0532-4Google Scholar
Crabtree, R.E. (1995). Chemical composition and energy content of deep-sea demersal fishes from tropical and temperate regions of the western North Atlantic. Bulletin of Marine Science 56 (2): 434449.Google Scholar
Crabtree, R.E. & Sulak, K.J. (1986). A contribution to the life history and distribution of Atlantic species of the deep-sea fish genus Conocara (Alepocephalidae). Deep-Sea Research Part A, Oceanographic Research Papers. 33: 11831201.CrossRefGoogle Scholar
Crabtree, R.E., Sulak, K.J., & Musick, J.A. (1985). Biology and distribution of the species of Polyacanthonotus (Pisces: Notacanthiformes) in the western North Atlantic. Bulletin of Marine Science 36:235248.Google Scholar
Craig, J., Jamieson, A.J., Hutson, R., Zuur, A.F., & Priede, I.G. (2010). Factors influencing the abundance of deep pelagic bioluminescent zooplankton in the Mediterranean Sea. Deep-Sea Research I 57: 14741484. doi:10.1016/j.dsr.2010.08.005CrossRefGoogle Scholar
Craig, J., Priede, I.G., Aguzzi, J., Company, J.B. & Jamieson, A.J. (2015). Abundant bioluminescent sources of low-light intensity in the deep Mediterranean Sea and North Atlantic Ocean. Marine Biology 162:16371649.Google Scholar
Craig, J., Youngbluth, M., Jamieson, A.J., & Priede, I.G. (2015). Near seafloor bioluminescence, macrozooplankton and macroparticles at the Mid-Atlantic Ridge. Deep-Sea Research I 98: 6275.CrossRefGoogle Scholar
Crane, J.M. (1968). Bioluminescence in the batfish Dibranchus atlanticus. Copeia 1968: 410411.Google Scholar
CSIRO Marine & Atmospheric Research. (2011). Deepsea Flathead, Hoplichthys haswelli. In: Fishes of Australia, accessed 16 October 2015, http://www.fishesofaustralia.net.au/home/species/3374Google Scholar
Currie, R.I. (1983). Research ship design & logistics. In: MacDonald, A.G. & Priede, I.G. (eds.) Experimental Design at Sea, pp. 387402. London: Academic Press.Google Scholar
Cuvier, G., Pietsch, T. (ed.). (1995). Historical Portrait of the Progress of Ichthyology, from Its Origins to Our Own Time. Baltimore, MD: Johns Hopkins.Google Scholar
Cuvier, G. & Vallenciennes, M. (1828). Histoire Naturelle des Poissons. Paris: Tome Premier. Levrault.Google Scholar
D’Onghia, G., Indennidate, A., Giove, A., Savini, A., Capezzuto, F., Sion, L., Vertino, A., & Maiorano, P. (2011). Distribution and behaviour of deep-sea benthopelagic fauna observed using towed cameras in the Santa Maria di Leuca cold-water coral province. Marine Ecology Progress Series 443: 95110.CrossRefGoogle Scholar
D’Ongia, G., Maiorano, P., & Sion, L. (2008) A review on the reproduction of Grenadiers in the Mediterranean with new data on the gonad Maturity and fecundity. American Fisheries Society Symposium 63: 169184.Google Scholar
D’Onghia, G., Politou, C.-Y., Bozzano, A., Lloris, D., Rotllant, G., Sion, L., & Mastrototaro, F. (2004). Deep-water fish assemblages in the Mediterranean Sea. Scientia Marina 68 (Suppl. 3): 8799.Google Scholar
D’Onghia, G., Sion, L., Maiorano, P., Mytilineou, C., Dalessandro, S., Carlucci, R., & Desantis, S. (2006). Population biology and life strategies of Chlorophthalmus agassizii Bonaparte, 1840 (Pisces: Osteichthyes) in the Mediterranean Sea. Marine Biology 149: 435446. doi:10.1007/s00227-005–0231-yGoogle Scholar
Dagit, D.D., Hareide, N., & Clò, S. (2007). Chimaera monstrosa. The IUCN Red List of Threatened Species. Version 2014.3. <www.iucnredlist.org>. Downloaded on 31 January 2015..+Downloaded+on+31+January+2015.>Google Scholar
Dahlhoff, E., Schneidemann, S., & Somero, G.N. (1990). Pressure-temperature interactions on M4-Lactate Dehydrogenases from hydrothermal vent fishes: Evidence for adaptation to elevated temperatures by the Zoarcid Thermarces andersoni, but not by the Bythitid, Bythites hollisi. Biological Bulletin 179: 134139.Google Scholar
Dalzell, P. & Pauly, D. (1989) Assessment of the fish resources of Southeast Asia, with emphasis on the Banda and Arafura seas. Netherlands Journal of Sea Research 24: 641650.Google Scholar
Danovaro, R., Company, J.B., Corinaldesi, C., D’Onghia, G., Galil, B.,et al. (2010). Deep-sea biodiversity in the Mediterranean Sea: The known, the unknown, and the unknowable. PLoS ONE 5(8): e11832. doi:10.1371/journal.pone.0011832CrossRefGoogle ScholarPubMed
Danovaro, R., Della, Croce N., Dell’Anno, A., & Pusceddu, A. (2003). A depocenter of organic matter at 7800m depth in the SE Pacific Ocean. Deep-Sea Research I 50: 14111420.CrossRefGoogle Scholar
Darwin, C. (1859). On the Origin of Species by Means of Natural Selection. London: Murray.Google Scholar
Das, M.K. & Nelson, J.S. (1996). Revision of the percophid genus Bembrops (Actinopterygii: Perciformes). Bulletin of Marine Science 59(1):944.Google Scholar
Davenport, J. & Kjørsvik, E. (1986). Buoyancy in the Lumpsucker Cyclopterus lumpus. Journal of the Marine Biological Association of the UK 66: 159174.CrossRefGoogle Scholar
Davesne, D., Friedman, M., Barriel, V., Lecointre, G., Janvier, P., Gallut, C., & Otero, O. (2014). Early fossils illuminate character evolution and interrelationships of Lampridiformes (Teleostei, Acanthomorpha). Zoological Journal of the Linnean Society 2014 (172): 475498.Google Scholar
Davies, A.J., Hosein, S., & Merrett, N.R. (2012). Haematozoans from deep water fishes trawled off the Cape Verde Islands and over the Porcupine Seabight, with a revision of species within the genus Desseria (Adeleorina: Haemogregarinidae). Folia Parasitologica 59 (1): 111.Google Scholar
Davies, A.J. & Merrett, N.R. (1998). Presumptive viral erythrocytic necrosis (VEN) in the benthopelagic fish Coryphaenoides (Nematonurus) armatus from the abyss west of Portugal. Journal of the Marine Biological Association of the UK 78: 10311034.CrossRefGoogle Scholar
Davies, I.E. & Barham, E.G. (1969). The Tucker opening-closing micronekton net and its performance in a study of the deep scattering layer. Marine Biology 2: 127131.Google Scholar
Davies, R., Cartwright, J., Pike, J., & Line, C. (2001). Early Oligocene initiation of North Atlantic Deep Water formation. Nature 410: 917920.CrossRefGoogle ScholarPubMed
Davis, M.P. (2015). Evolutionary relationships of the deep-sea pearleyes (Aulopiformes: Scopelarchidae) and a new genus of pearleye from Antarctic waters. Copeia, 2015(1): 6471.Google Scholar
Davis, M.P. & Fielitz, C. (2010). Estimating divergence times of lizardfishes and their allies (Euteleostei: Aulopiformes) and the timing of deep-sea adaptations. Molecular Phylogenetics and Evolution 57: 11941208.CrossRefGoogle ScholarPubMed
Davis, M.P., Holcroft, N.I., Wiley, E.O., Sparks, J.S., & Smith, W.L. (2014). Species-specific bioluminescence facilitates speciation in the deep sea. Marine Biology 161:11391148.Google Scholar
Davison, P.C., Checkley, D.M. Jr, Koslow, J.A., & Barlow, J. (2013). Carbon export mediated by mesopelagic fishes in the northeast Pacific Ocean. Progress in Oceanography 116: 1430.Google Scholar
Davison, P.C., Lara-Lopez, A., & Koslow, J.A. (2015). Mesopelagic fish biomass in the southern California current ecosystem. Deep-Sea Research II 112:129142.Google Scholar
De Leo, F.C., Smith, C.R., Rowden, A.A., Bowden, D.A., & Clark, M.R. (2010). Submarine canyons: hotspots of benthic biomass and productivity in the deep sea. Proceedings of the Royal Society of London Series B. 277:27832792. doi:10.1098/rspb.2010.0462Google Scholar
Deng, X., Wagner, H.-J. ,& Popper, A.N. (2011). The inner ear and its coupling to the swim bladder in the deep-sea fish Antimora rostrata (Teleostei: Moridae). Deep-Sea Research I 58: 2737. doi:10.1016/j.dsr.2010.11.001CrossRefGoogle Scholar
Denton, E.J. (1970). On the organization of reflecting surfaces in some marine animals. Philosophical Transactions of the Royal Society of London 258B: 285313.Google ScholarPubMed
Denton, E.J., Gilpin-Brown, J.B., & Wright, P.G. (1972). The angular distribution of the light produced by some mesopelagic fish in relation to their camouflage. Proceedings of the Royal Society B 182:145158. doi:10.1098/rspb.1972.0071Google Scholar
Denton, E.J. & Marshall, N.B. (1958). The buoyancy of bathypelagic fishes without a gas-filled swimbladder. Journal of the Marine Biological Association of the UK. 69: 409435.Google Scholar
Denton, J.E. & Yousef, M.K. (1976). Body composition and organ weights of rainbow trout, Salmo gairdneri. Journal of Fish Biology 8: 489499.Google Scholar
Desoutter, M. & Chapleau, F. (1997). Taxonomic status of Bathysolea profundicola and B. polli (Soleidae; Pleuronectiformes) with notes on the genus. Ichthyological Research 44(4): 399412.CrossRefGoogle Scholar
DeVaney, S.C., Hartel, K.E., & Themelis, D.E. (2009). The first records of Neocyema (Teleostei: Saccopharyngiformes) in the Western North Atlantic with comments on its relationship to Leptocephalus holti Schmidt 1909. Northeastern Naturalist 16(3): 409414.CrossRefGoogle Scholar
Devine, J.A., Baker, K.D., & Haedrich, R.L. (2006). Fisheries: Deep-sea fishes qualify as endangered. Nature 439: 29.Google Scholar
Diaz, H.F. & Bradley, R.S. (eds.). (2004). The Hadley Circulation: Present, Past and Future. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Doel, R.E., Levin, T.J., & Marker, M.K. (2006). Extending modern cartography to the ocean depths: military patronage, Cold War priorities, and the Heezen-Tharp mapping project, 1952–1959. Journal of Historical Geography 32, 605626.Google Scholar
Dolar, M.L., Walker, W.A., Kooyman, G.L., & Perrin, W.F. (2003). Comparative feeding ecology of Spinner dolphins (Stenella longirostris) and Fraser’s dolphins (Lagenodelphis hosei) in the Sulu Sea. Marine Mammal Science 19: 119.CrossRefGoogle Scholar
Dolgov, A.V., Drevetnyak, V., Sokolov, K.M., Grekov, A.A., & Shestopal, I.P. (2008). Biology and fisheries of roughead grenadier in the Barents Sea. American Fisheries Society Symposium 63: 343363.Google Scholar
Douglas, R.H., Partridge, J.C. & Marshall, N. J. (1998). The eyes of deep-sea fish I: Lens pigmentation, tapeta and visual pigments. Progress in Retinal and Eye Research. 17: 597636.CrossRefGoogle Scholar
Drazen, J.C. (2002). Energy budgets and feeding rates of Coryphaenoides acrolepis and C. armatus. Marine Biology 140: 677686.Google Scholar
Drazen, J.C. (2007). Depth related trends in proximate composition of demersal fishes in the eastern North Pacific. Deep-Sea Research I 54: 203219.Google Scholar
Drazen, J.C., Bailey, D.M., Ruhl, H.A., & Smith, K.L. Jr. (2012), The role of carrion supply in the abundance of deep water fish off California. PLoS ONE 7(11): e49332. doi:10.1371/journal.pone.0049332CrossRefGoogle ScholarPubMed
Drazen, J.C., Bird, L.E., & Barry, J.P. (2007). Development of a hyperbaric trap-respirometer for the capture and maintenance of live deep-sea organisms. Limnology & Oceanography: Methods 3: 488498.Google Scholar
Drazen, J.C., Buckley, T.W., & Hoff, G.R. (2001). The feeding habits of slope dwelling macrourid fishes in the eastern North Pacific. Deep-Sea Research I 48: 909995.CrossRefGoogle Scholar
Drazen, J.C., Dugan, B., & Friedman, J.R. (2013). Red muscle proportions and enzyme activities in deep-sea demersal fishes Journal of Fish Biology 83: 15921612. doi:10.1111/jfb.12268CrossRefGoogle ScholarPubMed
Drazen, J.C., Friedman, J.R., Condon, N.E., Aus, E.J., Gerringer, M.E., Keller, A.A., & Clarke, M.E. (2015). Enzyme activities of demersal fishes from the shelf to the abyssal plain. Deep-Sea Research I 100: 117126.CrossRefGoogle Scholar
Drazen, J.C. & Haedrich, R.L. (2012). A continuum of life histories in deep-sea demersal fishes. Deep-Sea Research I 61: 3442.Google Scholar
Drazen, J.C., Popp, B., & Choy, C. (2008). Bypassing the abyssal benthic food web: Macrourid diet in the eastern North Pacific inferred from stomach content and stable isotopes analyses. Limnology and Oceanography 53: 26442654.CrossRefGoogle Scholar
Drazen, J.C. & Robison, B.H. (2004). Direct observations of the association between a deep-sea fish and a giant scyphomedusa. Marine and Freshwater Behaviour and Physiology 37(3): 209214, doi: 10.1080/10236240400006190CrossRefGoogle Scholar
Drazen, J.C. & Seibel, B.A. (2007). Depth-related trends in metabolism of benthic and benthopelagic deep-sea fishes. Limnology &. Oceanography 52(5): 23062316.CrossRefGoogle Scholar
Drazen, J.C. & Sutton, T.T. (2016). Dining in the deep: The feeding ecology of deep-sea fishes. Annual Reviews in Marine Science 9: 337366.Google Scholar
Drazen, J.C. & Yeh, J. (2012). Respiration of four species of deep-sea demersal fishes measured in situ in the eastern North Pacific. Deep-Sea Research I 60:16.Google Scholar
Ducklow, H.W., Steinberg, D.K., & Buessler, K.O. (2001). Upper ocean carbon export and the biological pump. Oceanography 14: 5058.CrossRefGoogle Scholar
Duffy, C.A.J. (1997). Further records of the goblin shark, Mitsukurina owstoni (Lamniformes: Mitsukurinidae), from New Zealand. New Zealand Journal of Zoology 24: 2,167–171. doi: 10.1080/03014223.1997.9518111Google Scholar
Duhamel, G. (1995). Révision des genres Centriscops et Notopogon, Macroramphosidae des zones subtropicale et tempérée de l’hémisphére sud. Cybium 19(3): 261303.Google Scholar
Duhamel, G., Hulley, P.-A. Causse, R., Koubbi, P., Vacchi, M., Pruvost, P., et al. (2014). Chapter 7, Biogeographic patterns of fish. In: De Broyer, C., Koubbi, P., Griffiths, H.J., Raymond, B., Udekem, d’Acoz C. d’, et al. (eds.) Biogeographic Atlas of the Southern Ocean, pp. 328498. Cambridge: Scientific Committee on Antarctic Research.Google Scholar
Duhamel, G. & King, N. (2007). Deep-sea snailfish (Scorpaeniformes: Liparidae) of genera Careproctus and Paraliparis from the Crozet Basin (Southern Ocean). Cybium 31(3): 379387.Google Scholar
Dulčić, J. (2002). First record of scalloped ribbon fish, Zu cristatus (Pisces: Trachipteridae), eggs in the Adriatic Sea. Journal of Plankton Research. 24: 12451246.CrossRefGoogle Scholar
Duméril, A.M.C. (1806). Zoologie analytique, ou méthode naturelle de classification des animaux, rendue plus facile à l’aide de tableaux synoptiques. Paris: Allais.Google Scholar
Dunlap, P.V., Takami, M., Wakatsuki, S., Hendry, T.A., Sezaki, K., & Fukui, A. (2014). Inception of bioluminescent symbiosis in early developmental stages of the deep-sea fish, Coelorinchus kishinouyei (Gadiformes: Macrouridae). Ichthyological Research 61: 5967. doi: 10.1007/s10228-013–0374-7Google Scholar
Eastman, J.T., DeVries, A.L., Coalson, R.E., Nordquist, R.E., & Boyd, R.B. (1979). Renal conservation of antifreeze peptide in Antarctic eelpout, Rhigophila dearborni. Nature 282: 217221.CrossRefGoogle ScholarPubMed
Eastman, T., Hikida, R.S., & Devries, A.L. (1994). Buoyancy studies and microscopy of skin and subdermal extracellular matrix of the Antarctic snailfish, Paraliparis devriesi. Journal of Morphology 220: 85101.Google Scholar
Ebert, D.A. (2013). Deep–sea Cartilaginous Fishes of the Indian Ocean. Volume 1. Sharks. FAO Species Catalogue for Fishery Purposes. No. 8, Vol. 1. Rome: FAO.Google Scholar
Eddy, F.B. & Handy, R.D. (2012). Ecological and Environmental Physiology of Fishes. Oxford: Oxford University PressCrossRefGoogle Scholar
Ellerby, D.J. (2010). How efficient is a fish? Journal of Experimental Biology 213: 37653767. doi: 10.1242/jeb.034520Google Scholar
Ellis, J.E., Rowe, S., & Lotze, H.K. (2015). Expansion of hagfish fisheries in Atlantic Canada and worldwide. Fisheries Research 161: 2433.CrossRefGoogle Scholar
Enns, T., Scholander, P.F., & Bradstreet, E.D. (1965). Effect of hydrostatic pressure on gases dissolved in water. Journal of Physical Chemistry 69: 389391.CrossRefGoogle ScholarPubMed
Erdmann, M.V. (1999). An account of the first living coelacanth known to scientists from Indonesian waters. Environmental Biology of Fishes 54: 439443.Google Scholar
Erlandson, J,M,, Moss, M.L., & Des Lauriers, M. (2008). Life on the edge: early maritime cultures of the Pacific Coast of North America. Quaternary Science Reviews 27: 22322245.CrossRefGoogle Scholar
Eschmeyer, W. N. (ed.) (2014). Catalog of fishes: genera, species, references. (http://research.calacademy.org/research/ichthyology/catalog/fishcatmain.asp). Electronic version accessed 5 May 2014.Google Scholar
Espino, M., Castillo, R., & Fernández, F (1995). Biology and fisheries of the Peruvian Hake. In: Alheit, J. & Pitcher, T.J. (eds.) Hake: Biology, Fisheries and Markets, pp. 339363. London: Chapman & Hall.CrossRefGoogle Scholar
Etnoyer, P.J., Wood, J., & Shirley, T.C. (2010). How large is the seamount biome? Oceanography 23:206209.CrossRefGoogle Scholar
Evseenko, S.A. & Bol’shakova, Y.Y. (2014). First finding of a juvenile of the streamer fish Agrostichthys parkeri (Regalecidae) near the Walvis Ridge (Southern Atlantic). Journal of Ichthyology 54: 608610.Google Scholar
Evseenko, S.A. & Shtaut, M.I. (2005). On the species composition and distribution of ichthyoplankton and micronekton in the Costa Rica Dome and adjacent areas of the tropical Eastern Pacific. Journal of Ichthyology 45: 513525.Google Scholar
Eyring, C.F., Christensen, R.J., & Raitt, R.W. (1948). Reverberation in the sea. Journal of the Acoustical Society of America 20:462475.CrossRefGoogle Scholar
FAO. (2009). Report of the Technical Consultation on International Guidelines for the Management of Deep-sea Fisheries in the High Seas, Rome. 4–8 February and 25–29 August 2008, FAO Fisheries and Aquaculture Report, 881.Google Scholar
FAO. (2014). Deep-sea fisheries. www.fao.org/fishery/topic/4440/en accessed 18 June 2014.Google Scholar
FAO. (2015). Species Fact Sheets Reinhardtius hippoglossoides (Walbaum, 1792). www.fao.org/fishery/species/2544/en, Accessed 21 December 2015.Google Scholar
FAO. (2016). Species Fact Sheets Hexanchus griseus (Bonnaterre, 1788) www.fao.org/fishery/species/2003/en accessed 25 February 2016Google Scholar
Farias, I., Morales-Nin, B., Lorance, P., & Figueiredo, I. (2013). Black scabbardfish, Aphanopus carbo, in the northeast Atlantic: distribution and hypothetical migratory cycle. Aquatic Living Resources 26: 333342.Google Scholar
Farr, H.K. (1980). Multibeam bathymetric sonar: Sea beam and hydro chart. Marine Geodesy 4: 7793.CrossRefGoogle Scholar
Fasham, M.J.R. (ed.). (2003). Ocean Biogeochemistry, The Role of the Ocean Carbon Cycle in Global Change. Berlin: Springer.Google Scholar
Feagans-Bartow, J.N. & Sutton, T.T. (2014). Ecology of the oceanic rim: pelagic eels as key ecosystem components. Marine Ecology Progress Series 502: 257266.CrossRefGoogle Scholar
Feldman, G.C., Kuring, N.A., Ng, C., Esaias, W.E., Mcclain, C.R., Elrod, J.A., Maynard, N., Endres, D., Evands, R., Brown, J., Walsh, S., Carle, M. & Podesta, G. (1989). Ocean Color: Availability of the Global Data Set. Eos 70: 634641.Google Scholar
Fernandes, P.G. & Cook, R.M. (2013). Reversal of fish stock decline in the Northeast Atlantic. Current Biology 23: 14321437. http://dx.doi.org/10.1016/j.cub.2013.06.016CrossRefGoogle ScholarPubMed
Fernandez-Arcaya, U., Drazen, J.C., Murua, H., Ramirez-Llodra, E., Bahamon, N., Recasens, L., Rotllant, G., & Company, J.B. (2016). Bathymetric gradients of fecundity and egg size in fishes: a Mediterranean case study. Deep Sea Research I 116:106117. doi.org/10.1016/j.dsr.2016.08.005CrossRefGoogle Scholar
Fernholm, B. & Holmberg, K. (1975). The eyes in three genera of hagfish (Eptatretus, Paramyxine and Myxine) – a case of degenerative evolution. Vision Research 15: 253259.CrossRefGoogle ScholarPubMed
Figueiredo, I., Bordalo-Machado, P., Reis, S., Sena-Carvalho, D., Blasdale, T., Newton, A., & Gordo, L. S. (2003). Observations on the reproductive cycle of the black scabbardfish (Aphanopus carbo Lowe, 1839) in the NE Atlantic. ICES Journal of Marine Science 60: 774779.Google Scholar
Figueroa, D.E., Díaz de Astarloa, J.M., & Martos, P. (1998). Mesopelagic fish distribution in the southwest Atlantic in relation to water masses. Deep-Sea Research I 45: 317332.CrossRefGoogle Scholar
Fischer, W. & Bianchi, G. (1984). FAO Species identification sheets for fishery purposes, Western Indian Ocean, Fishing Area 51. Rome: Food and Agriculture Organization of the United Nations.Google Scholar
Fisher, J.A.D., Frank, K.T., Petrie, B. & Leggett, W.C. (2014) Life on the edge: environmental determinants of tilefish (Lopholatilus chamaeleonticeps) abundance since its virtual extinction in 1882. ICES Journal of Marine Science. doi: 10.1093/icesjms/fsu053.Google Scholar
Fleury, A.G. & Drazen, J.C. (2013). Abyssal scavenging communities attracted to Sargassum and fish in the Sargasso Sea. Deep Sea Research I 72: 141147. doi:10.1016/j.dsr.2012.11.004CrossRefGoogle Scholar
Flynn, A.J. & Marshall, N.J. (2013). Lanternfish (Myctophidae) zoogeography off Eastern Australia: a comparison with physicochemical biogeography. PLoS ONE 8(120): e80950. doi:10.1371/journal.pone.0080950Google Scholar
Fock, H., Uiblein, F., Köster, F., & von Westernhagen, H. (2002). Biodiversity and species-environment relationships of the demersal fish assemblage at the Great Meteor Seamount (subtropical NE Atlantic), sampled by different trawls. Marine Biology 141:185199. doi: 10.1007/s00227-002–0804-yGoogle Scholar
Follesa, M.C., Porcu, C., Cabiddu, S., Davini, M.A., Sabatini, A., & Cau, A. (2007). First observations on the reproduction of Alepocephalus rostratus Risso, 1820 (Osteichthyes, Alepocephalidae) from the Sardinian Channel (Central-Western Mediterranean). Marine Ecology 28 (Suppl. 1), 7581.Google Scholar
Foote, K.G. (1983). Linearity of fisheries acoustics, with addition theorems. Journal of the Acoustical Society of America 73: 19321940.CrossRefGoogle Scholar
Forbes, E. (1844). Report on the Mollusca and Radiata of the Aegean Sea, and on their distribution, considered as bearing on geology. Report of the British Association for the Advancement of Science. 13: 129193Google Scholar
Forbes, E. & Godwin-Austen, R. (1859). The Natural History of the European Seas. London: John Van Voorst.Google Scholar
Forman, J.S. & Dunn, M.R. (2010). The influence of ontogeny and environment on the diet of lookdown dory, Cyttus traversi. New Zealand Journal of Marine and Freshwater Research 44:4, 329342. doi: 10.1080/00288330.2010.523080CrossRefGoogle Scholar
Fornshell, J.A. & Tesei, A. (2013). The development of SONAR as a tool in marine biological research in the twentieth century. International Journal of Oceanography (2013), Article ID 678621. http://dx.doi.org/10.1155/2013/678621Google Scholar
Forster, G.R. (1964). Line-fishing on the continental slope. Journal of the Marine Biological Association of the UK 44, 277284.CrossRefGoogle Scholar
Forster, G.R. (1968). Line-fishing on the continental slope II. Journal of the Marine Biological Association of the UK 48, 479483.Google Scholar
Forster, G.R. (1973). Line fishing on the continental slope: the selective effect of different hook patterns. Journal of the Marine Biological Association of the UK 53, 749751.Google Scholar
Fossen, I. & Bergstad, O.A. (2006). Distribution and biology of blue hake, Antimora rostrata (Pisces: Moridae), along the mid-Atlantic Ridge and off Greenland. Fisheries Research 82: 1929.Google Scholar
Fossen, I., Cotton, C.F., Bergstad, O.A., & Dyb, J.E. (2008). Species composition and distribution patterns of fishes captured by longlines on the Mid-Atlantic Ridge. Deep-Sea Research II 55: 203217.CrossRefGoogle Scholar
Francis, M.P. & Duffy, C. (2002). Distribution, seasonal abundance and bycatch of basking sharks (Cetorhinus maximus) in New Zealand, with observations on their winter habitat. Marine Biology 140(4):831842.Google Scholar
Francis, M.P., Hurst, R.J., McArdle, B.H., Bagley, N.W., & Anderson, O.F. (2002). New Zealand demersal fish assemblages. Environmental Biology of Fishes 65: 215234.Google Scholar
Francis, R.I.C.C. & Clark, M.R. (2005). Sustainability issues for orange roughy fisheries. Bulletin of Marine Science 76(2): 337351.Google Scholar
Franco, M.A.L., Braga, A.C., Nunan, G.W.A., & Costa, P.A.S. (2009). Fishes of the family Ipnopidae (Teleostei: Aulopiformes) collected on the Brazilian continental slope between 11° and 23°S. Journal of Fish Biology 75: 797815.Google Scholar
Franz, V. (1907). Bau des Eulenauges und theorie des Teleskopauges. Biologisches Zentralblatt 27: 271278, 341–350.Google Scholar
Fraser, P.J., Cruickshank, S.F., & Shelmerdine, R.L. (2003). Hydrostatic pressure effects on vestibular hair cell afferents in fish and crustacea. Journal of Vestibular Research 13: 235242.Google Scholar
Fricke, H. (1997). Living coelacanths: values, eco-ethics and human responsibility. Marine Ecology Progress Series. 161:115.CrossRefGoogle Scholar
Fricke, H., Hissmann, K., Schauer, J., Erdmann, M., Moosa, M.K., & Plante, R. (2000). Biogeography of Indonesian coelacanths. Nature 403: 38.CrossRefGoogle ScholarPubMed
Fricke, H.W., Schauer, J., Hissmann, K., Kasang, L., & Plante, R. (1991). Coelacanth Latimeria chalumnae aggregates in caves: observations in their resting habitat and social behaviour. Environmental Biology of Fishes 30: 281285.CrossRefGoogle Scholar
Fricke, R. (1992). Revision of the family Draconettidae (Teleostei), with descriptions of two new species and a new subspecies. Journal of Natural History 26(1):165195.CrossRefGoogle Scholar
Fricke, R. (2010). Centrodraco atrifilum, a new deepwater dragonet species from eastern Australia (Teleostei: Draconettidae). Stuttgarter Beiträge zur Naturkunde A, Neue Serie 3: 341346.Google Scholar
Friedman, J.F., Condon, N.E., & Drazen, J.C. (2012). Gill surface area and metabolic enzyme activities of demersal fishes associated with the oxygen minimum zone off California. Limnology & Oceanography 57(6): 17011710.Google Scholar
Friedman, M. & Sallan, L.C. (2012). Five hundred million years of extinction and recovery: a phanerozoic survey of Large-scale diversity patterns in fishes. Palaeontology 55: 707742.CrossRefGoogle Scholar
Froese, R. (2014) The making of FishBase. http://www.fishbase.de/manual/English/fishbasethe_making_of_fishbase.htm (accessed 5 May 2014)Google Scholar
Froese, R. & Pauly, D. (eds.). (2016). FishBase. World Wide Web electronic publication. www.fishbase.org, version (01/2016).Google Scholar
Froese, R. & Sampang, A. (2004). Taxonomy and biology of seamount fishes. In: Morato, T. & Pauly, D. (eds.) Seamounts: Biodiversity and Fisheries, pp. 2531. Fisheries Centre Research Reports. Vancouver: University of British Columbia 12(5).Google Scholar
Fudge, D. S., Levy, N., Chiu, S., & Gosline, J. M. (2005). Composition, morphology and mechanics of hagfish slime. Journal of Experimental Biology 208: 46134625.CrossRefGoogle ScholarPubMed
Fujikura, K., Kojima, S., Tamaki, K, Maki, Y., Hunt, J., & Okutan, T. (1999). The deepest chemosynthesis-based community yet discovered from the hadal zone, 7326 m deep, in the Japan Trench. Marine Ecology Progress Series 190: 1726.CrossRefGoogle Scholar
Fukui, A. & Kitagawa, Y. (2006). Dolichopteryx rostrata, a new species of spookfish (Argentinoidea: Opisthoproctidae) from the eastern North Atlantic Ocean. Ichthyological Research 53: 712.CrossRefGoogle Scholar
Gage, J.D. & Tyler, P. (1991). Deep-Sea Biology: A Natural History of Organisms at the Deep-Sea Floor. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Gaither, M.R., Bowen, B.W., Rocha, L.A., & Briggs, J.C. (2016). Fishes that rule the world: circumtropical distributions revisited. Fish and Fisheries 17: 664679.CrossRefGoogle Scholar
Gallo, N.D., Cameron, J., Hardy, K., Fryer, P., Bartlett, D.H., & Levin, L.A. (2015). Submersible- and lander-observed community patterns in the Mariana and New Britain Trenches: Influence of productivity and depth on benthic community structure epibenthic and scavenging communities. Deep-Sea Research I 99: 119133.Google Scholar
Garcia, D.M., Rice, J., & Charles, A. (eds.). (2014). Governance of Marine Fisheries and Biodiversity Conservation: Interaction and Co-evolution, Oxford: Wiley-Blackwell.CrossRefGoogle Scholar
Garcıa, V.B., Lucifora, L.O. & Myers, R.A. (2008). The importance of habitat and life history to extinction risk in sharks, skates, rays and chimaeras. Proceedings of the Royal. Society B 275: 8389.CrossRefGoogle ScholarPubMed
Garcia-Castellanos, D., Estrada, F., Jiménez-Munt, I., Gorini, C., Fernàndez, M., Vergés, J., & De Vicente, R. (2009). Catastrophic flood of the Mediterranean after the Messinian salinity crisis. Nature 462: 778782.CrossRefGoogle ScholarPubMed
García-Mederos, A.M., Tuset, V.M., Santana, J.I., & González, J.A. (2010). Reproduction, growth and feeding habits of stout beardfish Polymixia nobilis (Polymixiidae) off the Canary Islands (NE Atlantic). Journal of Applied Ichthyology 26: 872880.CrossRefGoogle Scholar
Garman, S.H. (1899). The fishes. Reports on an exploration off the west coast of Mexico, Central America, South America, and off the Galapagos Islands, in charge of Alexander Agassiz, by the U.S. Fish Commission steamer ‘Albatross’, during 1891, Lieut. Commander Z. L. Tanner, U.S.N. commanding. Pt. 26. Memoirs of the Museum of Comparative Zoology 24: 1421.Google Scholar
Gartner, J.V., Crabtree, R.E., & Sulak, K.J. (1997). Feeding at depth. In: Randall, D.J., & Farrell, A.P. (eds.) Deep-Sea Fishes, pp. 115–93. San Diego: Academic Press.Google Scholar
Gaskett, A.C., Bulman, C., He, X., & Goldsworthy, S.D. (2001). Diet composition and guild structure of mesopelagic and bathypelagic fishes near Macquarie Island, Australia. New Zealand Journal of Marine and Freshwater Research 35(3): 469476. doi: 10.1080/00288330.2001.9517016CrossRefGoogle Scholar
Gebbie, G. & Huybers, P. (2012). The mean age of ocean waters inferred from radiocarbon observations: sensitivity to surface sources and accounting for mixing histories. Journal of Physical Oceanography 42: 291305. doi: http://dx.doi.org/10.1175/JPO-D-11-043.1CrossRefGoogle Scholar
GEBCO. (2014). General bathymetric chart of the oceans. Accessed 3 June 2014. www.gebco.net/data_and_products/gridded_bathymetry_data/Google Scholar
Gebruk, A.V., Chevaldonné, P., Shank, T., Lutz, R.A., & Vrijenhoek, R.C. (2000). Deep-sea hydrothermal vent communities of the Logatchev area (14°45’N Mid-Atlantic Ridge): diverse biotopes and high biomass. Journal of the Marine Biological Association of the U.K. 80: 383393.CrossRefGoogle Scholar
Geistdoerfer, P. (1991). Ichtyofaune associée à l’hydrothermalisme océanique et description de Thermobiotes mytiligeiton, nouveau genre et nouvelle espèce de Synaphobranchidae (Pisces, anguilliformes) de l’océan Pacifique. Comptes rendus de l’Académie des sciences. Série 3, Sciences de la vie 312(3): 9197.Google Scholar
Geistdoerfer, P. (1994a) Careproctus hyaleius, nouvelle espèce de poisson cyclopteridae (Liparinae) de l’écosystème hydrothermal de la dorsale de l’Océan Pacifique oriental (13°N). Cybium 18(3): 325333.Google Scholar
Geistdoerfer, P. (1994b) Pachycara thermophilum, une nouvelle espèce de poisson Zoarcidae des sites hydrothermaux de la dorsale Médio-Atlantique. Cybium 18(2): 109115.Google Scholar
Geistdoerfer, P. (1999). Thermarces pelophilum, espece nouvelle de Zoarcidae associee a l’emission de fluides froids au niveau du prisme d’accretion de la barbade, Ocean Atlantique Nord-Ouest. Cybium 23(1): 511.Google Scholar
Gessner, C. (1555). Historiae animalium Vol 3 Bk 4. De piscium et aquatilium animantium natura. Zurich: Christoph Froschoverus.Google Scholar
Gibbs, A.G. (1997). Biochemistry at depth. In Randall, D.J. & Farrell, A.P. (eds.) Deep-Sea Fishes. San Diego: Academic Press.Google Scholar
Gibbs, A. & Somero, G. N. (1990). Na+-K+-adenosine triphosphatase activities in gills of marine teleost fishes: changes with depth, size and locomotory activity level. Marine Biology 106: 315321.CrossRefGoogle Scholar
Gibbs, R.H., Clarke, T.A., & Gomon, J.R. (1983). Taxonomy and Distribution of the Stomioid Fish Genus Eustomias (Melanostomiidae), I: Subgenus Nominostomias. Washington, DC: Smithsonian Contributions to Zoology, no. 380.Google Scholar
Gilbert, C.H. (1905). The deep-sea fishes. Pt 2. In: Jordan, D.S. & Everyman, B.W. (eds.) The Aquatic Resources of the Hawaiian Islands, pp. 577713. Bulletin of the U.S. Fisheries Commission. 23(2)Google Scholar
Gilbert, C.H. & Hubbs, C.L. (1928). The macrourid fishes of the Philippine Islands and the East Indies. U.S. National Museum Bulletin 100, 1(pt. 7): 369588.Google Scholar
Gilchrist, J. D. F. (1922). Deep-sea fishes procured by the S.S. “Pickle” (Part I). Report of the Fisheries and Marine Biological Survey, Union of South Africa (Rep. 2, art. 3): 41–79.Google Scholar
Gillibrand, E.J.V., Bagley, P.M., Jamieson, A., Herring, P.J., Partridge, J.C., Collins, M.A., Milne, R., & Priede, I.G. (2007). Deep sea benthic bioluminescence at artificial food falls, 1000 to 4800m depth, in the Porcupine Seabight and Abyssal Plain. North East Atlantic Ocean. Marine Biology 150: 10531060. doi :10.1007/s00227-006–0407-0.Google Scholar
Gillibrand, E.J.V., Jamieson, A.J., Bagley, P.M., Zuur, A.F., & Priede, I.G. (2007). Seasonal development of a deep pelagic bioluminescent layer in the temperate Northeast Atlantic Ocean. Marine Ecology Progress Series 341: 3744Google Scholar
Gilmore, R.G., (1997). Lipogramma robinsi, a new basslet from the tropical western Atlantic, with descriptive and distributional notes on L. flavescens and L. anabantoides (Perciformes: Grammatidae). Bulletin of Marine Science 60(3): 782788.Google Scholar
Gjøsæter, J. (1979). Mesopelagic fish. In: Saetre, R. & Silva, R.P. (eds.) The Marine Fish Resources of Mozambique, pp. 101114. Bergen: Institute of marine Research.Google Scholar
Gjøsaeter, J. (1984). Mesopelagic fish, a large potential resource in the Arabian Sea. Deep-Sea Research. Part A 31: 10191035. On line publication www.fao.org/WAIRDOCS/FNS/FN130E/ch11.htm accessed 2 October 2016.CrossRefGoogle Scholar
Gjøsaeter, J. & Kawaguchi, K. (1980). A review of the world resources of Mesopelagic Fish, Food and Agriculture Organization of the United Nations. FAO Fisheries Technical Paper 193.Google Scholar
Glover, A.G., Higgs, N., & Horton, T. (2014). World register of deep-sea species. Accessed at www.marinespecies.org/deepsea on 2014-05-0Google Scholar
Godbold, J.A., Bailey, D.M. Collins, M.A., Gordon, J.D.M., Spallek, W.A., & Priede, I.G. (2013). Putative fishery-induced changes in biomass and population size structures of demersal deep-sea fishes in ICES Sub-area VII, North East Atlantic Ocean. Biogeosciences 10: 529539. doi:10.5194/bg-10–529-2013, 2013Google Scholar
Gomez, C., Williams, A.J., Nicol, S.J., Mellin, C., Loeun, K.L., & Bradshaw, C.J.A. (2015). Species distribution models of tropical deep-sea snappers. PLoS ONE 10 (6): e0127395. doi:10.1371/journal.pone.0127395CrossRefGoogle ScholarPubMed
González-Costas, F. & Murua, H. (2008). An analytical assessment of the Roughhead grenadier stock in NAFO subareas 2 and 3. American Fisheries Society Symposium 63: 319342.Google Scholar
Goode, G.B. & Bean, T.H. (1895). Oceanic Ichthyology, a treatise on the Deep-Sea and Pelagic Fishes of the world based chiefly upon collections made by the steamers Blake, Albatross and Fish Hawk in the Northwestern Atlantic Ocean. Smithsonian Institution, Special Bulletin 2. Washington, DC: Government Printing Office.Google Scholar
Gordeeva, N.V. (2013). Genetic divergence in the tribe Electronini (Myctophidae). Journal of Ichthyology 53: 575584.CrossRefGoogle Scholar
Godø, O.R., Huse, I., & Michalsen, K. (1997). Bait defence behaviour of wolffish and its impact on long-line catch rates. ICES Journal of Marine Science 54: 273275.Google Scholar
Gordon, J.D.M. (2003). The Rockall Trough, Northeast Atlantic: the cradle of deep-sea biological oceanography that is now being subject to unsustainable fishing activity. Journal of Northwest Atlantic Fisheries Science 31: 5783.CrossRefGoogle Scholar
Gore, M.A., Rowat, D., Hall, J., Gell, F.R., & Ormond, R.F. (2008). Transatlantic migration and deep mid-ocean diving by basking shark. Biology Letters 4: 395398.Google Scholar
Govoni, J.J., Olney, J.E., Markle, D.F., & Curtsinger, W.R. (1984). Observations on structure and evaluation of possible functions of the vexillum in larval Carapidae (Ophidiiformes). Bulletin of Marine Science 34: 6070.Google Scholar
Graae, M. J. F. (1967). Lestidium bigelowi, a new species of paralepidid fish with photophores. Breviora 277:110.Google Scholar
Grant, S.M. (2006). An exploratory fishing survey and biological resource assessment of Atlantic hagfish (Myxine glutinosa) occurring on the southwest slope of the Newfoundland Grand Bank. Journal of Northwest Atlantic Fishery Science. 36: 91110.CrossRefGoogle Scholar
Grassle, J.F. (2000). The Ocean Biogeographic Information System (OBIS): an on-line, worldwide atlas for accessing, modeling and mapping marine biological data in a multidimensional geographic context. Oceanography 13: 57.CrossRefGoogle Scholar
Grassle, J.F., Sanders, H.L., Hessler, R.R., Rowe, G.T., & MacLennan, T. (1975). Pattern and zonation: a study of the bathyal megafauna using the research submersible Alvin. Deep-Sea Research 22: 643659.Google Scholar
Gray, E.W. (1789). Observations on the class of animals called, by Linnaeus, Amphibia; particularly on the means of distinguishing those serpents which are venomous, from those which are not so. Philosophical Transactions of the Royal Society of London 79: 2136.Google Scholar
Greer-Walker, M. & Pull, G. A. (1975). A survey of red and white muscle in marine fish. Journal of Fish Biology 7: 295300.Google Scholar
Greer-Walker, M., Santer, R.M., Benjamin, M., & Norman, D. (1985). Heart structure of some deep-sea fish (Teleostei: Macrouridae). Journal of Zoology London (A) 205: 7589.CrossRefGoogle Scholar
Gregory, S., Brown, J., & Belchier, M. (2014). Ecology and distribution of the grey notothen, Lepidonotothen squamifrons, around South Georgia and Shag Rocks, Southern Ocean. Antarctic Science 26: 239249. doi:10.1017/S0954102013000667CrossRefGoogle Scholar
Greven, H., Walker, Y., & Zanger, K. (2009). On the structure of teeth in the viperfish Chauliodus sloani Bloch & Schieder, 1801 (Stomiidae). Bulletin of Fish Biology 11: 8798.Google Scholar
Grogan, D. & Lund, R. (2004). The origin and relationships of early Chondrichthyes. In: Carrier, J.C., Musick, J.A., & Heithanus, M.R. (eds.) Biology of Sharks and Their Relatives, pp. 331. London: CRC Press.Google Scholar
Guinot, G, Adnet, S., Cavin, L., & Cappetta, H. (2013). Cretaceous stem chondrichthyans survived the end-Permian mass extinction. Nature Communications 4: 2669. doi: 10.1038/ncomms3669CrossRefGoogle ScholarPubMed
Günther, A.C.L.G. (1880). An introduction to the study of fishes. Edinburgh: Black.Google Scholar
Günther, A.C.L.G. (1887). Report on the Deep-Sea Fishes collected by the H.M.S. Challenger during the years, 1873–1876. Report on the Scientific Results of the Voyage of the HMS Challenger during the years 1873–1876 22.Google Scholar
Haedrich, R. L. (1964). Food habits and young stages of North Atlantic Alepisaurus. Breviora 201: 115.Google Scholar
Haedrich, R.L. (1965). Identification of a deep-sea mooring-cable biter. Deep-Sea Research 12: 773776.Google Scholar
Haedrich, R.L. (1977). A sea lamprey from the deep ocean. Copeia 1977: 767768.CrossRefGoogle Scholar
Haedrich, R.L. & Craddock, J.E. (1968). Distribution and biology of the opisthoproctid fish Winteria telescopa Brauer 1901. Breviora 294: 111.Google Scholar
Haedrich, R.L. & Merrett, N.R. (1988). Summary atlas of deep-living demersal fishes in the North Atlantic Basin. Journal of Natural History 22(5): 13251362.CrossRefGoogle Scholar
Haedrich, R.L. & Merrett, N.R. (1990). Little evidence for faunal zonation or communities in the deep sea demersal fish faunas. Progress in Oceanography 24: 239250.Google Scholar
Haffner, R.E. (1952). Zoogeography of the Bathypelagic Fish, Chauliodus. Systematic Zoology 1: 112133.Google Scholar
Halliday, R.G., Themelis, D.E., & Hickey, W.M. (2012). Demersal fishes caught with bottom gillnets and baited gears at 500–2 800 m on the Continental Slope off Nova Scotia, Canada. Journal of Northwest Atlantic Fisheries Science 44: 3140.CrossRefGoogle Scholar
Hamon, N., Sepulchre, O., Lefebvre, V. & Ramstein, G. (2013). The role of eastern Tethys seaway closure in the Middle Miocene Climatic Transition (ca. 14 Ma). Climate of the Past 9: 26872702. www.clim-past.net/9/2687/2013/ doi:10.5194/cp-9-2687-2013Google Scholar
Hansen, C. A. & Sidell, B. D. (1983). Atlantic hagfish cardiac-muscle metabolic basis of tolerance to anoxia. American Journal of Physiology 244: R356R362.Google ScholarPubMed
Hansen, J., Sato, M., Hearty, P., Ruedy, R., Kelley, M., Masson-Delmotte, V., Russell, G., Tselioudis, G., Cao, J., Rignot, E., Velicogna, I., Tormey, B., Donovan, B., Kandiano, E., von Schuckmann, K., Kharecha, P., Legrande, A.N., Bauer, M., & Lo, K.-W. (2016). Ice melt, sea level rise and superstorms: evidence from paleoclimate data, climate modeling, and modern observations that 2°C global warming could be dangerous. Atmospheric Chemistry and Physics 16: 37613812.CrossRefGoogle Scholar
Harbison, G.R. & Janssen, J. (1987). Encounters with a Swordfish (Xiphias gladius) and Sharptail Mola (Masturus lanceolatus) at Depths Greater Than 600 Meters. Copeia 1987 (2): 511513.Google Scholar
Harbison, G.R., Smith, K.L., & Backus, R.H. (1973). Stygiomedusa fabulosa from the North Atlantic, its taxonomy, with a note on its natural history. Journal of the Marine Biological Association of the UK 53: 615617.CrossRefGoogle Scholar
Hare, J.A. & Marancik, (2005). Psychrolutidae: tadpole or fathead sculpins. In: Richards, W.J. (ed,) Early Stages of Atlantic Fishes: An Identification Guide for the Western Central North Atlantic, pp. 11911192. Boca Raton, FL: CRC Press.Google Scholar
Harold, A.S (1994). A taxonomic revision of the Sternoptychid genus Polyipnus (Teleostei: Stomiiformes) with an analysis of phylogenetic relationships. Bulletin of Marine Science 54(2): 428534.Google Scholar
Harold, A.S., Hartel, K.E., Craddock, J.E., & Moore, J.A. (2002). Hatchetfishes and relatives. Family Sternoptychidae. In: Collette, B.B. & Klein-MacPhee, G. (eds.) Bigelow and Schroeder’s Fishes of the Gulf of Maine, pp. 184190. Washington, DC: Smithsonian Institution Press.Google Scholar
Harold, A.S. & Lancaster, K. (2003). A new species of hatchetfish genus Argyripnus (Stomiiformes: Sternoptychidae) from the Indo-Pacific. Proceedings of the Biological Society of Washington 116: 883891.Google Scholar
Harper, A. A., Macdonald, A. G., Wardle, C. S. & Pennec, J.-P. (1987). The pressure tolerance of deep-sea fish axons: results of Challenger cruise 6B/85. Comparative Biochemistry and Physiology 88A: 647–53.Google Scholar
Harrison, C.M.H. (1966). On the first halosaur leptocephalus: from Madeira. Bulletin of the British Museum (Natural History) Zoology 14: 445486.Google Scholar
Harrison, C.M.H. & Palmer, G. (1968). On the neotype of Radiicephalus elongatus Osorio with remarks on its biology. Bulletin of the British Museum (Natural History) Zoology 16(5): 185208.Google Scholar
Hartman, O. & Emery, K. O. (1956). Bathypelagic coelenterates. Limnology and Oceanography 1: 304312.Google Scholar
Hawkins, A.D. & Amorim, M.C.P. (2000). Spawning sounds of the male haddock, Melanogrammus aeglefinus. Environmental Biology of Fishes 59: 2941.CrossRefGoogle Scholar
Hayes, A.J. & Sim, A. J. W. (2011). Ratfish (Chimaera) spine injuries in fishermen. Scottish Medical Journal 56: 161163. doi: 10.1258/smj.2011.011115Google Scholar
Hayes, D.B., Ferreri, C.P., & Taylor, W.W. (2013). Active Fish Capture Methods. In Zale, A.V., Parrish, D.L., & Sutton, T.M. (eds.) Fisheries Techniques, 3rd edition, Chapter 7. Bethesda, MD: American Fisheries Society.Google Scholar
Haygood, M.G. (1993). Light organ symbioses in fishes, Critical Reviews in Microbiology 19(4): 191216. doi: 10.3109/10408419309113529CrossRefGoogle ScholarPubMed
Hebert, P.D.N., Cywinska, A., Ball, S.L., & deWaard, J.R. (2003), Biological identifications through DNA barcodes. Proceedings of the Royal Society London B 270: 313321. doi:10.1098/rspb.2002.2218CrossRefGoogle ScholarPubMed
Hector, J. (1875). Descriptions of five new species of fishes obtained in the New-Zealand seas by H. M. S -“Challenger” Expedition, July 1874. Annals and Magazine of Natural History (Series 4) 15(85) (art. 11): 7882.Google Scholar
Heezen, B.C. (1969). The world rift system: an introduction to the symposium. Technophysics 8: 269279.Google Scholar
Heezen, B.C. & Hollister, C.D. (1971). Face of the deep. New York: Oxford University Press.Google Scholar
Heger, A., King, N., Wigham, B.D., Jamieson, A.J., Bagley, P.M., Allan, L., Pfannkuche, O., & Priede, I.G. (2007). Benthic Bioluminescence in the bathyal North East Atlantic: luminescent responses of Vargula norvegica (Ostracoda: Myodocopida) to predation by the deep water eel (Synaphobranchus kaupii). Marine Biology 151 (4): 14711478. doi:10.1007/s00227-006–0587-7Google Scholar
Heimberg, A.M., Cowper-Sallari, R., S´emon, M., Donoghue, P.C.J., & Peterson, K.J. (2010). MicroRNAs reveal the interrelationships of hagfish, lampreys, and gnathostomes and the nature of the ancestral vertebrate. Proceedings of the National Academy of Sciences of the USA 107: 1937919383. doi: 10.1073/pnas.1010350107CrossRefGoogle ScholarPubMed
Heinke, F. (1913). Untersuchungen über die Scholle-Generalbericht I Schollen-fischerei und Schonmassregeln. Vorlaeufige kurze Euebersicht über die wichtigsten Ergebnisse de Berichts. Rapports et Procès-Verbaux des Réunions du Conseil Permanent pour L’exploration de la Mer 180: 170.Google Scholar
Helfman, G.S., Collette, B.B., Facey, D.E., & Bowen, B.W. (2009). The Diversity of fishes, Biology Evolution and Ecology, 2nd edition. Oxford: Wiley-Blackwell.Google Scholar
Henriques, C. (2004). In situ lander observations of deep-sea fishes in the Eastern Atlantic Ocean. PhD thesis University of Aberdeen.Google Scholar
Henry, L.-A., Stehmann, M.F.W., De Clippele, L., Findlay, H.S., Golding, N., & Roberts, J.M. (2016). Seamount egg-laying grounds of the deep-water skate Bathyraja richardsoni. Journal of Fish Biology. doi:10.1111/jfb.13041CrossRefGoogle Scholar
Herring, P. (2002). The Biology of the Deep Ocean. Oxford: Oxford University Press.Google Scholar
Herring, P.J. (1971). Bioluminescence in an evermannellid fish. Journal of Zoology (London) 181: 297307.Google Scholar
Herring, P.J. (1987). Systematic distribution of bioluminescence in living organisms. Journal of Bioluminescence and Chemiluminescence 1: 147163.CrossRefGoogle ScholarPubMed
Herring, P.J. (1992). Bioluminescence of the oceanic apogonid fishes Howella brodiei and Florenciella lugubris. Journal of the Marine Biological Association of the UK 72: 139148. doi:10.1017/S0025315400048840Google Scholar
Herring, P.J. (2000). Species abundance, sexual encounter and bioluminescent signalling in the deep sea. Philosophical Transactions: Biological Sciences 355: 12731276.Google Scholar
Herring, P.J. & Cope, C. (2005). Red bioluminescence in fishes: on the suborbital photophores of Malacosteus, Pachystomias and Aristostomias. Marine Biology 148: 383394. doi: 10.1007/s00227-005–0085-3Google Scholar
Herring, P.J., Gaten, E., & Shelton, P.M.J. (1999). Are vent shrimps blinded by science? Nature 398: 116.Google Scholar
Hirakawa, N., Suzuki, N., Narimatsu, Y., Saruwatari, T., & Ohno, A. (2007). The spawning and settlement season of Chlorophthalmus albatrossis along the Pacific Coast of Japan. The Raffles Bulletin of Zoology 14: 167170.Google Scholar
Hissmann, K., Fricke, H., & Schauer, J. (1998). Patterns of time and space utilization in coelacanths (Latimeria chalumnae) determined by ultrasonic telemetry. Marine Biology 136: 943952.CrossRefGoogle Scholar
Hissmann, K., Fricke, H., & Schauer, J. (2008). Population monitoring of the Coelacanth (Latimeria chalumnae). Conservation Biology 12(4): 759765.CrossRefGoogle Scholar
Ho, H.-C. (2013). Two new species of the batfish genus Malthopsis (Lophiiformes: Ogcocephalidae) from the Western Indian Ocean. Zootaxa 3716: 289300.CrossRefGoogle ScholarPubMed
Ho, H.-C. & Last, P.R. (2013). Two new species of the coffinfish genus Chaunax (Lophiiformes: Chaunacidae) from the Indian Ocean. Zootaxa 3710(5): 436448.CrossRefGoogle ScholarPubMed
Ho, H.-C. & McGrouther, M. (2015). A new anglerfish from eastern Australia and New Caledonia (Lophiiformes: Chaunacidae: Chaunacops), with new data and submersible observation of Chaunacops melanostomus. Journal of Fish Biology 86(3): 940951.CrossRefGoogle ScholarPubMed
Ho, H.-C., Roberts, C.D., & Shao, K.-T. (2013). Revision of batfishes (Lophiiformes: Ogcocephalidae) of New Zealand and adjacent waters, with description of two new species of the genus Malthopsis. Zootaxa 3626(1): 188200.CrossRefGoogle ScholarPubMed
Ho, H.-C., Roberts, C. D., & Stewart, A. L. (2013). A review of the anglerfish genus Chaunax (Lophiiformes: Chaunacidae) from New Zealand and adjacent waters, with descriptions of four new species. Zootaxa 3620(1): 89111.CrossRefGoogle ScholarPubMed
Hochachka, P.W., Storey, K.B., & Baldwin, J. (1975). Gill citrate synthase from an Abyssal fish. Comparative Biochemistry & Physiology 52B: 4349.Google ScholarPubMed
Høines, Å. S. & Gundersen, A. C. (2008). Rebuilding the stock of Northeast Arctic Greenland halibut (Reinhardtius hippoglossoides). Journal of Northwest Atlantic Fisheries Science 41: 107117. doi:10.2960/J.v41.m618CrossRefGoogle Scholar
Holdsworth, E.W.H. (1874). Deep-Sea Fishing and Fishing Boats. London: Stanford.Google Scholar
Honjo, S., Manganini, S.J., Krishfield, R.A., & Francois, R. (2008). Particulate organic carbon fluxes to the ocean interior and factors controlling the biological pump: a synthesis of global sediment trap programs since 1983. Progress in Oceanography 76: 217285.CrossRefGoogle Scholar
Hopkins, T.L. & Baird, R.C. (1973). Diet of the Hatchetfish Sternoptyx diaphana. Marine Biology 21: 3446.CrossRefGoogle Scholar
Hopkins, T.L. & Baird, R.C. (1981). Trophodynamics of the Fish Valenciennellus tripunctulatus. I. Vertical Distribution, Diet and Feeding Chronology. Marine Ecology Progress Series 5: 110.Google Scholar
Horn, P.L. (2003). Stock structure of bluenose (Hyperoglyphe antarctica) off the north-east coast of New Zealand based on the results of a detachable hook tagging programme. New Zealand Journal of Marine and Freshwater Research 37(3): 623631. doi:10.1080/00288330.2003.9517193CrossRefGoogle Scholar
Horn, P.L., Dunn, M.R., & Forman, J. (2013). The diet and trophic niche of orange perch, Lepidoperca aurantia (Serranidae: Anthiinae) on Chatham Rise, New Zealand. Journal of Ichthyology 53: 310316.Google Scholar
Hort, J. (1911). The ‘Michael Sars’ North Atlantic Deep-Sea Expedition, 1910. The Geographical Journal 37: 349377.Google Scholar
Hovis, W.A., Clark, D.K., Anderson, F., Austin, R.W., Wilson, W.H., Baker, E.T., Ball, D., Gordon, H.R., Mueller, J.L., El-Sayed, S.Z., Sturm, B., Wrigley, R.C., & Yentsch, C.S. (1980). Nimbus-7 Coastal Zone Color Scanner: system description and initial imagery. Science 210: 6063.CrossRefGoogle ScholarPubMed
Howes, G.J. (1991). Biogeography of gadoid fishes. Journal of Biogeography 18: 595622.CrossRefGoogle Scholar
Hsü, K.J., Ryan, W.B.F. & Cita, M.B. (1973). Late Miocene desiccation of the Mediterranean. Nature 242: 240244. doi:10.1038/242240a0Google Scholar
Hubbs, C. L. (1935). Half mile down by William Beebe Review. Copeia 1935 (2): 105.CrossRefGoogle Scholar
Hubbs, C. L. & Iwamoto, T. (1977). A new genus (Mesobius), and three new bathypelagic species of Macrouridae (Pisces, Gadiformes) from the Pacific Ocean. Proceedings of the California Academy of Sciences (Series 4), 41 (7): 233251.Google Scholar
Hughes, G.M. (1966). The dimensions of fish gills in relation to their function. Journal of Experimental Biology 45: 177195.Google Scholar
Hughes, G.M. (1979). Morphometry of fish gas exchange organs in relations to their respiratory function. In: Ali, M.A. (ed.) Environmental Physiology of Fishes, pp. 3346. New York: Plenum Press.Google Scholar
Hughes, G.M. & Iwai, T. (1978). A morphometric study of the gills in some Pacific deep-sea fishes. Journal of Zoology 184: 155170, doi:10.1111/j.1469–.7998.1978.tb03272.xGoogle Scholar
Hughes, S.E. (1981). Initial U.S. Exploration of Nine Gulf of Alaska Seamounts and Their Associated Fish and Shellfish Resources. Marine Fisheries Review. 42: 2633.Google Scholar
Hulley, P.A. & Lutjeharms, J.R.E. (1995). The south-western limit for the warm-water, mesopelagic ichthyofauna of the Indo-West Pacific: lanternfish (Myctophidae) as a case study. South African Journal of Marine Science 15(1): 185205. doi: 10.2989/02577619509504843CrossRefGoogle Scholar
Hulley, P.A. & Prosch, R.M. (1987). Mesopelagic fish derivatives in the southern Benguela upwelling region South African Journal of Marine Science 5(1): 597611, doi: 10.2989/025776187784522289CrossRefGoogle Scholar
Humphreys, R.L. Jr., Winans, G.A., & Tagami, D.T. (1989). Synonymy and life history of the North Pacific pelagic armorhead, Pseudopentaceros wheeleri Hardy (Pisces: Pentacerotidae). Copeia 1989(1): 142153.Google Scholar
ICES. (2012). Report of the Working Group on the Biology and Assessment of Deep-Sea Fisheries Resources (WGDEEP), 28 March–5 April, Copenhagen, Denmark: ICES, CM 2012/ACOM:17.Google Scholar
ICES. (2013). Manual for the Midwater Ring Net sampling during IBTS Q1: The International Bottom Trawl Survey Working Group. Copenhagen. Denmark: International Council for the Exploration of the Sea.Google Scholar
ICES. (2014). International Council for Exploration of the Sea. Report of the Report of the Working Group on Widely Distributed Stocks. (WGWIDE), 26 August–1 September 2014, ICES Headquarters. Copenhagen, Denmark: ICES CM 2014/ACOM:15.Google Scholar
ICES. (2015). Report of the Working Group on Biology and Assessment of Deep-Sea Fisheries Resources (WGDEEP), 20–17 March 2015. Copenhagen, Denmark: ICES, CM 2015/ACOM:17.Google Scholar
ICES Advice. (2015). ICES Advice on fishing opportunities, catch, and effort Iceland Sea and Oceanic Northeast Atlantic Ecoregions. 2.3.4a Beaked redfish (Sebastes mentella) in Subareas V, XII, and XIV (Iceland and Faroes grounds, north of Azores, east of Greenland) and NAFO Subareas 1+2 (deep pelagic stock > 500 m). Accessed online 8 April 2016. www.ices.dk/sites/pub/Publication%20Reports/Advice/2015/2015/smn-dp.pdf+500+m).+Accessed+online+8+April+2016.+www.ices.dk/sites/pub/Publication%20Reports/Advice/2015/2015/smn-dp.pdf>Google Scholar
ICES.NWWG. (2015). Report of the North-Western Working Group (NWWG), 28 April–5 May, ICES HQ. Copenhagen Denmark: ICES, CM 2015/ACOM:07.Google Scholar
Ichino, M.C., Clark, M.R., Drazen, J.C., Jamieson, A., Jones, D.O.B., Martin, A.P., Rowden, A.A., Shank, T.M., Yancey, P.H., & Ruhl, H.A. (2015). The distribution of benthic biomass in hadal trenches: a modelling approach to investigate the effect of vertical and lateral organic matter transport to the seafloor. Deep-Sea Research I 100: 2133.CrossRefGoogle Scholar
IGBP, IOC, SCOR. (2013). Ocean Acidification Summary for Policymakers – Third Symposium on the Ocean in a High-CO2 World. Stockholm, Sweden: International Geosphere-Biosphere Programme.Google Scholar
Iglésias, S.P., Dettai, A., & Ozouf-Costaz, C. (2012). Barbapellis pterygalces, new genus and new species of a singular eelpout (Zoarcidae: Teleostei) from the Antarctic deep waters. Polar Biology 35(2): 215220.Google Scholar
Iglésias, S.P., Toulhout, L. & Sellos, D.P. (2010). Taxonomic confusion and market mislabelling of threatened skates: important consequences for their conservation status. Aquatic Conservation 20: 319333.CrossRefGoogle Scholar
Ikeda, T. (1996). Metabolism, body composition, and energy budget of the mesopelagic fish Maurolicus muelleri in the Sea of Japan. Fishery Bulletin 94: 4958.Google Scholar
Ikeda, T. & Hirakawa, K. (1998). Metabolism and body composition of zooplankton in the cold mesopelagic zone of the southern Japan Sea. Plankton Biology and Ecology 45(1): 3144.Google Scholar
Inoue, J. G., Miya, M., Lam, K., Tay, B. H., Danks, J. A., Bell, J., Walker, T. I., & Venkatesh, B. (2010). Evolutionary origin and phylogeny of the modern holocephalans (Chondrichthyes: Chimaeriformes): a mitogenomic perspective. Molecular Biology and Evolution 27: 25762586.Google Scholar
Inoue, J.G., Miya, M., Miller, M.J., Sado, T., Hanel, R., Hatooka, K., Aoyama, J., Minegishi, Y., Nishida, M., & Tsukamoto, K. (2010). Deep-ocean origin of the freshwater eels. Biology Letters 6: 363366. doi:10.1098/rsbl.2009.0989Google Scholar
International Commission on Zoological Nomenclature. (1999). International Code of Zoological Nomenclature, 4th edition. Singapore: International Commission on Zoological Nomenclature. Online version: http://iczn.org/iczn Accessed 24 June 2014.Google Scholar
Irigoien, X., Klevjer, A., Røstad, A., Martinez, U., Boyra, G., Acuña, J.L., Bode, A., Echevarria, F., Gonzalez-Gordillo, J.I., Hernandez-Leon, S., Agusti, S., Aksnes, D.L., Duarte, C.M., & Kaartvedt, S. (2014). Large mesopelagic fishes biomass and trophic efficiency in the open ocean. Nature Communications 5: 3271. doi: 10.1038/ncomms4271Google Scholar
Isaacs, J.D. & Schick, G.B. (1960). Deep-sea free instrument vehicle. Deep-Sea Research 7: 6167.Google Scholar
Isaacs, J.D. & Schwartzlose, R.A. (1975). Active animals of the deep-sea floor. Scientific American 233(4): 8491. doi: 10.1038/scientificamerican1075-84Google Scholar
Ivanov, V.V., Shapiro, G.I., Huthnance, J.M., Aleynik, D.L., & Golovin, P.N. (2004). Cascades of dense water around the world ocean. Progress in Oceanography 60: 4798.Google Scholar
Iwamoto, T. (1975). The abyssal fish Antimora rostrata (Günther). Comparative Biochemistry and Physiology Part B: Comparative Biochemistry 52: 711.CrossRefGoogle ScholarPubMed
Iwamoto, T. (2008a). A brief taxonomic history of grenadiers. American Fisheries Society Symposium 63: 313.Google Scholar
Iwamoto, T. (2008b). A note on the correct spelling of Coelorinchus (Caelorinchus) and C.coelorhincus (caelorhincus). American Fisheries Society Symposium 63: xi.Google Scholar
Iwamoto, T. (2015). Hoplostethus cadenati. The IUCN Red List of Threatened Species 2015: e.T21110117A21907969. http://dx.doi.org/10.2305/IUCN.UK.2015-4.RLTS.T21110117A21907969.en. Downloaded on 21 March 2017.CrossRefGoogle Scholar
Iwamoto, T. & Arai, T. (1987). A new grenadier Malacocephalus okamurai (Pisces: Gadiformes: Macrouridae) from the Western Atlantic. Copeia 1987 (1): 204208.Google Scholar
Iwamoto, T. & Orlov, A. (2006). Paracetonurus flagellicauda (Keofoed, 1927) (Macrouridae, Gadiformes, Teleostei), new records from the Mid-Atlantic Ridge and Madagascar Plateau. Proceedings of the California Academy of Sciences 4th Series 57(11): 379386.Google Scholar
Iwamoto, T. & Orlov, A. (2008). First Atlantic Record of Asthenomacrurus victoris Sazonov & Shcherbachev (Macrouridae, Gadiformes, Teleostei). Proceedings of the California Academy of Sciences, Series 4, 59(4): 125131.Google Scholar
Iwamoto, T. & Ungaro, N. (2002). A new grenadier (Gadiformes, Macrouridae) from the Mediterranean. Cybium 26(1): 2732.Google Scholar
Iwamoto, T. & Sazonov, Y.I. (1994). Revision of the genus Kumba (Pisces, Gadiformes, Macrouridae), with the description of three new species. Proceedings of the California Academy of Sciences 48(11): 221237.Google Scholar
Jackson, T.L. (2002). Preliminary guide to the identification of the early life history stages of giganturid fishes of the Western Central North Atlantic. NOAA Technical Memorandum NMFS-SEFSC-484.Google Scholar
Jacobs, D.K. & Lindberg, D.R. (1998). Oxygen and evolutionary patterns in the sea: Onshore/offshore trends and recent recruitment of deep-sea faunas. Proceedings of the National Academy of Sciences USA 95: 93969401.CrossRefGoogle ScholarPubMed
James, P.S.B.R. (1981). Exploited and potential capture fishery resources in the inshore waters of India. Central Marine Fisheries Research Institute Bulletin 30A: 7282. http://eprints.cmfri.org.in/726/1/James_psbr_72.pdf accessed on line 27 August 2015.Google Scholar
Jamieson, A.J. (2015). The Hadal Zone: Life in the Deepest Oceans. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Jamieson, A.J., Fujii, T., Mayor, D.J., Solan, M., & Priede, I.G. (2010). Hadal trenches: the ecology of the deepest places on Earth. Trends in Ecology & Evolution 25: 190197.Google Scholar
Jamieson, A.J., Fujii, T., Solan, M., Matsumoto, A.K., Bagley, P.M., & Priede, I.G. (2009). Liparid and macrourid fishes of the hadal zone: in situ observations of activity and feeding behaviour. Proceedings of the Royal Society of London B. 276: 10371045. doi:10.1098/rspb.2008.1670Google ScholarPubMed
Jamieson, A.J., Godø, O.R., Bagley, P.M., Partridge, J.C., & Priede, I.G. (2006). Illumination of trawl gear by mechanically stimulated bioluminescence. Fisheries Research 81: 276282.Google Scholar
Jamieson, A.J., Lacey, N.C., Lörz, A.-N., Rowden, A.A., & Piertney, S.B. (2013). The supergiant amphipod Alicella gigantea (Crustacea: Alicellidae) from hadal depths in the Kermadec Trench, SW Pacific Ocean. Deep-Sea Research II 92: 107113.CrossRefGoogle Scholar
Jamieson, A.J., Priede, I.G.,& Craig, J. (2012). Distinguishing between the abyssal macrourids Coryphaenoides yaquinae and C. armatus from in situ photography. Deep Sea Research I 64: 7885.Google Scholar
Jamieson, A.J., Boorman, B., & Jones, D.O.B. (2013). Deep-sea sampling. In: Eleftheriou, A. (ed.) Methods for the Study of Marine Benthos, pp. 285347, 4th edition. Oxford: Wiley-Blackwell.CrossRefGoogle Scholar
Jamieson, A.J. & Yancey, P.H. (2012). On the validity of the trieste flatfish: dispelling the myth. Biological Bulletin 222: 171175.Google Scholar
Janßen, F., Treude, T., & Witte, U. (2000). Scavenger assemblages under differing trophic conditions: a case study in the deep Arabian Sea. Deep Sea Research II 47: 29993026.CrossRefGoogle Scholar
Janssen, J. (2004). Lateral line sensory ecology. In Von der Emde, G., Mogdans, J., & Kapoor, B.G. (eds.) The Senses of Fish; Adaptations for the Reception of Natural Stimuli, pp. 231264. Berlin: Springer.Google Scholar
Janssen, J., Gibbs, R.H. Jr., & Pugh, P.R. (1989). Association of Caristius Sp. (Pisces: Caristiidae) with a Siphonophore, Bathyphysa conifer. Copeia 1989: 198201.CrossRefGoogle Scholar
Janssen, J. & Harbison, G.R. (1981). Fish in salps: the association of squaretails (Tetragonurus spp.) with pelagic tunicates. Journal of the Marine Biological Association of the UK 61(4): 917927.CrossRefGoogle Scholar
Janvier, P. (2010). MicroRNAs revive old views about jawless vertebrate divergence and evolution. Proceedings of the National Academy of Sciences of the USA 107: 1913719138.Google Scholar
Jeffreys, R.M., Lavaleye, M.S.S., Bergman, M.J.N., Duineveld, G.C.A., Witbaard, R., & Linley, T. (2010). Deep-sea macrourid fishes scavenge on plant material: Evidence from in situ observations. Deep-Sea Research I 57: 621627.Google Scholar
Jennings, S., Reynolds, J.D., Mills, S.C. (1998). Life history correlates of responses to fisheries exploitation. Proceedings of the Royal Society of London B 265: 333339.Google Scholar
Johannessen, A. & Monstad, T. (2003). Distribution, growth and exploitation of Greater Silver Smelt (Argentina silus, (Ascanius, 1775) in Norwegian waters. Journal of North Atlantic Fisheries Science 31: 319332.CrossRefGoogle Scholar
Johnson, G.D., Paxton, J.R., Sutton, T.T., Satoh, T.P., Sado, T., Nishida, M., & Miya, M. (2009). Deep-sea mystery solved: astonishing larval transformations and extreme sexual dimorphism unite three fish families. Biology Letters 5: 235239.Google Scholar
Johnson, J.Y. (1862). Descriptions of some new genera and species of fish obtained at Madiera. Proceedings of the Zoological Society of London 1862: 167180.Google Scholar
Johnson, J.Y. (1863). Descriptions of five new species of fishes obtained at Madeira. Proceedings of the Zoological Society of London 1863(33): 3646.Google Scholar
Johnson, M.W. (1948). Sound as a tool in marine ecology, from data on biological noises and the deep scattering layer. Journal of Marine Research 7(3): 443458.Google Scholar
Johnson, R.K. & Bertelsen, E. (1991). The fishes of the family Giganturidae: systematics, development, distribution and aspects of biology. Dana Reports 91: 145.Google Scholar
Johnson, R.K. (1974). Five new species and a new genus of alepisauroid fishes of the Scopelarchidae (Pisces: Myctophiformes). Copeia 1974(2): 449457.Google Scholar
Johnston, I.A. & Herring, P.J. (1985). The transformation of muscle into bioluminescent tissue in the fish Benthabella infans Zugmayer. Proceedings of the Royal Society of London B 255: 213218.Google Scholar
Jones, D.O.B., Bett, B.J., Wynn, R.B., & Masson, D.G. (2009). The use of towed camera platforms in deep-water science. Underwater Technology 28(2): 4150.CrossRefGoogle Scholar
Jones, E.C. (1971). Isistius brasiliensis, a squaloid shark, the probable cause of crater wounds on fishes and cetaceans. Fisheries Bulletin 69(4): 791798.Google Scholar
Jones, E.G. (1999). ‘Burial at sea’. Consumption and dispersal of large fish and cetacean food-falls by deep-sea scavengers in the Northeast Atlantic Ocean and the Eastern Mediterranean Sea. PhD thesis University of Aberdeen.Google Scholar
Jones, E.G., Collins, M.A., Bagley, P.M., Addison, S., & Priede, I.G. (1998). The fate of cetacean carcasses in the deep sea: observations on consumption rates and succession of scavenging species in the abyssal north-east Atlantic Ocean. Proceedings of the Royal Society of London. B 265: 11191127.Google Scholar
Jones, E.G., Tselepides, A., Bagley, P.M., Collins, M.A., & Priede, I.G. (2003). Bathymetric distribution of some benthic and benthopelagic species attracted to baited cameras and traps in the Eastern Mediterranean. Marine Ecology Progress Series 251: 7586.CrossRefGoogle Scholar
Jones, M.R.L. & Breen, B.B. (2013) Food and feeding relationships of three sympatric slickhead species (Pisces: Alepocephalidae) from northeastern Chatham Rise, New Zealand. Deep-Sea Research I 79: 19.Google Scholar
Jordan, D.S. (1917–1920). The genera of fishes from Linnaeus to Cuvier, 1758–1833, seventy five years with the accepted type of each. A contribution to the stability of scientific nomenclature. Stanford, CA: Stanford University.Google Scholar
Jordan, D.S. (1919). The Genera of Fishes, Part III from Guenther to Gill, 1859–1880, Twenty Two Years, with the Accepted Type of Each. Stanford, CA: Stanford University.Google Scholar
Jordan, D.S. & Starks, E.C. (1904). List of fishes dredged by the steamer Albatross off the coast of Japan in the summer of 1900, with descriptions of new species and a review of the Japanese Macrouridae. Bulletin of the U.S. Fish Commission 22[1902]: 577630, Pls. 1–8.Google Scholar
Jørgensen, J.M., Lomholt, J.P., Weber, R.E., & Malte, H. (eds.). (1998). The Biology of Hagfishes. London: Chapman & HallCrossRefGoogle Scholar
Jumper, G. Y. & Baird, R.C. (1991). Location by olfaction: a model and application to the mating problem in the deep sea hatchetfish Argyropelecus hemigyrnnus. The American Naturalist 138: 14311458.Google Scholar
Kaartvedt, S., Staby, A., & Aksnes, D.L. (2012). Efficient trawl avoidance by mesopelagic fishes causes large underestimation of their biomass. Marine Ecology Progress Series 456: 16. doi: 10.3354/meps09785Google Scholar
Kabasakal, H. (2006). Distribution and biology of the bluntnose sixgill shark Hexanchus griseus (Bonnaterre, 1788) (Chondrichthyes, Hexanchidae), from Turkish waters. Annales Series Historia Naturalis 16(1): 2936.Google Scholar
Kaiser, M.J. & Spencer, B.E. (1994) Fish scavenging in recently trawled areas. Marine Ecology Progress Series. 112: 4149.Google Scholar
Kanwisher, J. & Ebeling, A. (1957). Composition of the swimbladder gas in bathypelagic fishes. Deep-Sea Research 4(3): 211217.Google Scholar
Karmovskaya, E.S. & Merrett, N.R. (1998). Taxonomy of the deep-sea eel genus, Histiobranchus (Synaphobranchidae, Anguilliformes), with notes on the ecology of H. bathybius in the eastern North Atlantic. Journal of Fish Biology 53: 10151037.Google Scholar
Karmovskaya, E.S. & Parin, N.V. (1999). [ref. 24869] A new species of the genus Ilyophis (Synaphobranchidae, Anguilliformes) from the hydrothermal area Broken Spur (Mid-Atlantic Submarine Ridge). Voprosy Ikhtiologii 39(3): 316325.Google Scholar
Karuppasamy, P.K., Muraleedharan, K.R., Dineshkumar, P.K., & Nair, M. (2010). Distribution of mesopelagic micronekton in the Arabian Sea during the winter monsoon. Indian Journal of Marine Sciences 39: 227237.Google Scholar
Katsanevakis, S., Maravelias, C. D., & Kell, L. T. (2010). Landings profiles and potential métiers in Greek set longliners. ICES Journal of Marine Science, 67: 646656.Google Scholar
Kaufmann, R.S., Wakefield, W.W., & Genin, A. (1989). Distribution of epibenthic megafauna and lebensspuren on two central North Pacific seamounts. Deep-Sea Research 36: 18631896.CrossRefGoogle Scholar
Kawaguchi, K. & Butler, J.L. (1984). Fishes of the genus Nansenia (Microsomatidae) with descriptions of seven new species. Los Angeles County Museum Contributions in Science 352: 122.Google Scholar
Kawai, T. (2013). Revision of the peristediid genus Satyrichthys (Actinopterygii: Teleostei) with the description of a new species, S. milleri sp. nov. Zootaxa 3635(4): 419438.CrossRefGoogle ScholarPubMed
Kawai, T., Amaoka, K., & Serét, (2010). A new righteye flounder, Poecilopsetta multiradiata (Teleostei: Pleuronectiformes: Poecilopsettidae), from New Zealand and New Caledonia (South-West Pacific). Ichthylogical Research 57: 193198.CrossRefGoogle Scholar
Keeling, R.F., Körtzinger, A., & Gruber, N. (2010). Ocean deoxygenation in a warming world. Annual Review of Marine Science 2: 199229.CrossRefGoogle Scholar
Kemp, K.M., Jamieson, A.J., Bagley, P.M., McGrath, H., Bailey, D.M., Collins, M.A., & Priede, I.G. (2006). Consumption of large bathyal food fall, a six month study in the NE Atlantic. Marine Ecology-Progress Series 310: 6576.CrossRefGoogle Scholar
Kenaley, C.P. (2007). Revision of the stoplight loosejaw genus Malacosteus (Teleostei: Stomiidae: Malacosteinae), with description of a new species from the temperate Southern Hemisphere and Indian Ocean. Copeia 2007(4): 886900.CrossRefGoogle Scholar
Kenaley, C.P. (2009). Revision of Indo-Pacific species of the loosejaw dragonfish genus Photostomias (Teleostei: Stomiidae: Malacosteinae). Copeia 2009(1): 175189.CrossRefGoogle Scholar
Kenaley, C.P., DeVaney, S.C., & Fjeran, T.T. (2013). The complex evolutionary history of seeing red: molecular phylogeny and the evolution of an adaptive visual system in deep-sea dragonfishes (Stomiiformes: Stomiidae). Evolution 68–4: 9961013.CrossRefGoogle Scholar
Kim, S.-Y. (2012). Phylogenetic systematics of the family Pentacerotidae (Actinopterygii: Order Perciformes). Zootaxa 3366: 1111.Google Scholar
Kimura, S., Kohno, Y., Tsukamoto, Y., & Okiyama, M. (1990). Record of the Parabrotulid fish Parabrotula plagiophthalma from Japan. Japanese Journal of Ichthyology 37: 318320.Google Scholar
King, N.J., Bagley, P.M., & Priede, I.G. (2006). Depth zonation and latitudinal distribution of deep sea scavenging demersal fishes of the Mid-Atlantic Ridge, 42°–53°N. Marine Ecology Progress Series 319: 263274.Google Scholar
King, N.J., Bailey, D.M., & Priede, I.G. (2007). Introduction; role of scavengers in marine ecosystems. Marine Ecology Progress Series 350: 175178. doi: 10.3354/meps07186CrossRefGoogle Scholar
King, N.J. & Priede, I.G. (2008). Coryphaenoides armatus, the abyssal grenadier: distribution, abundance, and ecology as determined by baited landers. American Fisheries Society Symposium 63: 139161.Google Scholar
Kinzer, J., Böttger-Schnack, R., & Schulz, K. (1993). Aspects of horizontal distribution and diet of myctophid fish in the Arabian Sea with reference to the deep water oxygen deficiency. Deep-Sea Research II 40: 783800.Google Scholar
Kious, J.W. & Tilling, R. (1996). This Dynamic Earth: The Story of Plate Tectonics. Online booklet. Washington, DC: U.S. Geological Survey. http://pubs.usgs.gov/publications/text/dynamic.html Accessed 23 September 2014.Google Scholar
Kjerstad, M., Fossen, I., & Willemsen, H.M. (2003). Utilization of deep-sea sharks at Hatton Bank in the N Atlantic. Journal of NorthWest Atlantic Fishery Science 31: 333338.CrossRefGoogle Scholar
Klepadlo, C. (2011). Three new species of the genus Photonectes (Teleostei: Stomiiformes: Stomiidae: Melanostomiinae) from the Pacific Ocean. Copeia 2011(2): 201210.CrossRefGoogle Scholar
Klimpel, S., Busch, M.W., Kellermanns, E., Kleinertz, S., & Palm, H.W. (2009). Metazoan Deep-Sea Fish Parasites. Solingen,Germany: Verlag Natur & Wissenschaft.Google Scholar
Kloser, R.J., Sutton, C., Krusic-Golub, K., & Ryan, T.E. (2015). Indicators of recovery for orange roughy (Hoplostethus atlanticus) in eastern Australian waters fished from 1987. Fisheries Research 167(2015): 225235.CrossRefGoogle Scholar
Klug, S. & Kriwet, J. (2010). Timing of deep-sea adaptation in dogfish sharks: insights from a supertree of extinct and extant taxa. Zoologica Scripta 39: 331342.CrossRefGoogle Scholar
Knudsen, S.W., Møller, P.R., & Gravlund, P. (2007). Phylogeny of the snailfishes (Teleostei: Liparidae) based on molecular and morphological data. Molecular Phylogenetics and Evolution 44(2007): 649666.Google Scholar
Knudsen, S., Nielsen, J., & Uiblein, F. (2015). Spectrunculus grandis. The IUCN Red List of Threatened Species 2015: e.T18139045A60799962. http://dx.doi.org/10.2305/IUCN.UK.2015-4.RLTS.T18139045A60799962.en. Downloaded on 21 March 2017.CrossRefGoogle Scholar
Kobylianskii, S.G. (2006). Bathylagus niger sp. nova (Bathylagidae, Salmoniformes) a new species of Bathylagus from the subpolar waters of the Southern Ocean. Journal of Ichthyology 46(6): 413417.Google Scholar
Kobyliansky, S.G., Orlov, A.M., & Gordeeva, N.V. (2010). Composition of deep-sea pelagic ichthyocenes of the Southern Atlantic, from waters of the area of the Mid-Atlantic and Walvis Ridges. Journal of Ichthyology 50(10): 932949.CrossRefGoogle Scholar
Kock, K.-H. (1992). Antarctic fish and fisheries. Cambridge: Cambridge University Press.Google Scholar
Kooistra, W.H.C.F. & Medlin, L.K. (1996). Evolution of the diatoms (Bacillariophyta). IV A reconstruction of their age from small subunit rRNA coding regions and the fossil record. Molecular Phylogenetics and Evolution (Academic Press) 6(3): 391407. doi:10.1006/mpev.1996.0088Google Scholar
Koslow, A.J., Davison, P., Lara-Lopez, A., & Ohman, M.D. (2014). Epipelagic and mesopelagic fishes in the southern California Current System: ecological interactions and oceanographic influences on their abundance. Journal of Marine Systems 138: 2028.Google Scholar
Koslow, J.A. (1993) Community structure in North Atlantic deep-sea fishes. Progress in Oceanography 31: 321338.CrossRefGoogle Scholar
Koslow, J.A. (2007). The Silent Deep; the Discovery, Ecology and Conservation of the Deep-Sea. Chicago: University of Chicago Press.Google Scholar
Koslow, J.A., Boehlert, G.W., Gordon, J.D.M., Haedrich, R.L., Lorance, P., & Parin, N. (2000). Continental slope and deep-sea fisheries: implications for a fragile ecosystem. ICES Journal of Marine Science 57: 548557.CrossRefGoogle Scholar
Koslow, J.A., Bulman, C.M., & Lyle, J.M. (1994). The mid-slope demersal fish community off southeastern Australia. Deep-Sea Research I 41: 113141.Google Scholar
Koslow, J.A., Kloser, R.J., & Williams, A. (1997). Pelagic biomass and community structure over the mid-continental slope off southeastern Australia based upon acoustic and midwater trawl sampling. Marine Ecology Progress Series 146: 2135.Google Scholar
Kotlyar, A.N. (2008). Revision of the genus Poromitra (Melamphaeidae): Part 2. New species of the group P. crassiceps. Journal of Ichthyology 48(8): 581592.Google Scholar
Kotlyar, A.N. (2009). Revision of the genus Poromitra (Melamphaidae): Part 4. Species of P. cristiceps group: P. atlantica, P. oscitans, and P. agofonovae Kotlyar, species nova. Journal of Ichthyology 49(8): 563574.CrossRefGoogle Scholar
Kotrschal, K., Van Staaden, M.J., & Huber, R. (1998). Fish brains: evolution and environmental relationships. Reviews in Fish Biology and Fisheries 8: 373408.CrossRefGoogle Scholar
Krefft, G. (1976). Distribution patterns of oceanic fishes in the Atlantic Ocean. Revue Des Travaux De l’Institut Des Pêches Maritimes 40(3/4): 439460.Google Scholar
Krefft, G. (1990). Chimaeridae. In: Quero, J.C., Hureau, J.C., Karrer, C., Post, A., & Saldanha, L. (eds.) Check-List of the Fishes of the Eastern Tropical Atlantic (CLOFETA), pp. 111113. Lisbon: JNICT; Paris: SEI; and Paris: UNESCO. Vol. 1.Google Scholar
Kriwet, J. (2003). Lancetfish teeth (Neoteostei, Alepisauroidei) from the early cretaceous of Alcaine, Spain. Lethaia 36: 323332.CrossRefGoogle Scholar
Kriwet, J. & Benton, M.J. (2004). Neoselachian (Chondrichthyes, Elasmobranchii) diversity across the Cretaceous–Tertiary boundary. Palaeogeography, Palaeoclimatology, Palaeoecology 214: 181194.Google Scholar
Kriwet, J., Kiessling, W., & Klug, S. (2009). Diversification trajectories and evolutionary life-history traits in early sharks and batoids. Proceedings of the Royal Society of London B 276: 945951. doi:10.1098/rspb.2008.1441Google ScholarPubMed
Krueger, W.H. & Gibbs, R.H. (1966). Growth changes and sexual dimorphism in the stomiatoid fish Echiostoma barbatum. Copeia 1966(1): 4349.CrossRefGoogle Scholar
Kubota, T., Shiobara, Y., & Kubodera, T. (1991). Food habits of the frilled shark Chlamydoselachus anguineus collected from Suruga bay, central Japan. Nippon Suisan Gakkaishi 57(1): 1520. doi:10.2331/suisan.57.15CrossRefGoogle Scholar
Kukuev, E.I., Parin, N.V., & Trunov, I.A. (2013). Materials for the revision of the family Caristiidae (Perciformes): 3. Manefishes (genus Caristius) from moderate warm waters of the Pacific and Atlantic oceans with a description of three new species from the southeast Atlantic (C. barsukovi sp. n., C. litvinovi sp. n., C. walvisensis sp. n.). Journal of Ichthyology 53(8): 541561.Google Scholar
Kulczykowska, E, Popek, W., & Kapoor, B.G. (eds.). (2010). Biological Clock in Fish. Boca Raton, FL: CRC Press.CrossRefGoogle Scholar
Kulka, D., Hood, C., & Huntington, J. (2007). Recovery strategy for northern wolffish (Anarhichas denticulatus) and spotted wolffish (Anarhichas minor), and management plan for Atlantic wolffish (Anarhichas lupus) in Canada. Ottawa, Canada: Department of Fisheries and Oceans.Google Scholar
Kunzig, R. (2003) Deep-sea biology: living with the endless frontier. Science 302: 991.Google Scholar
Kunzmann, A. & Zimmerman, C. (1992). Aethotaxis mitopteryx, a high-Antarctic fish with benthopelagic mode of life. Marine Ecology Progress Series 88: 3340.CrossRefGoogle Scholar
Kuraku, S. & Kuratani, S. (2006). Timescale for cyclostome evolution inferred with a phylogenetic diagnosis of hagfish and lamprey cDNA sequences. Zoological Science 23: 10531064.CrossRefGoogle Scholar
Kusukawa, S. (2000). The Historia piscium (1986). Notes and Records of the Royal Society of London 54(2): 179197.Google Scholar
Kyne, P.M. & Simpfendorfer, C.A. (2007). A collation and summarization of available data on deepwater chondrichthyans: biodiversity, life history and fisheries. Burnaby, BC: IUCN SSC Shark Specialist Group for the Marine Conservation Biology Institute. At: www.flmnh.ufl.edu/fish/organizations/SSG/SSG.htmGoogle Scholar
Lack, M. (2008). Continuing CCAMLR’s Fight against IUU Fishing for Toothfish. Sydney: WWF Australia and TRAFFIC International.Google Scholar
La Mesa, M., Eastman, J.T., & Licandro, P. (2007). Feeding habits of Bathydraco marri (Pisces, Notothenioidei, Bathydraconidae) from the Ross Sea, Antarctica. Polar Biology 30: 541547. doi: 10.1007/s00300-006–0211-9CrossRefGoogle Scholar
Lampitt, R.S., Bett, B.J., Kiriakoulakis, K., Popova, E.E., Ragueneau, O., Vangriesheim, A., & Wolff, G.A. (2001). Material supply to the abyssal seafloor in the Northeast Atlantic. Progress in Oceanography 50: 2763.CrossRefGoogle Scholar
Lampitt, R.S. & Burnham, M.P. (1983). A free fall time lapse camera and current meter system “Bathysnap” with notes on the foraging behaviour of a bathyal decapod shrimp. Deep-Sea Research 30: 10091017.CrossRefGoogle Scholar
Lampitt, R.S., Merrett, N.R., & Thurston, M.H. (1983). Inter-relations of necrophagous amphipods, a fish predator, and tidal currents in the deep sea. Marine Biology 74: 7378.Google Scholar
Land, M.F. & Osorio, D.C. (2011). Marine optics: dark disguise. Current Biology 21, R918–R92. doi: 10.1016/j.cub.2011.10.009Google Scholar
Laptikhovsky, V. (2010). Migrations and structure of the species range in ridge-scaled rattail Macrourus carinatus (Southwest Atlantic) and their application to fisheries management. ICES Journal of Marine Science doi:10.1093/icesjms/fsq081CrossRefGoogle Scholar
Laptikhovsky, V., Arkhipin, A., & Brickle, P. (2008). Biology and distribution of grenadiers of the family Macrouridae around the Falkland Islands. American Fisheries Society Symposium 63: 261284.Google Scholar
Lauerman, L.M.L. & Kaufmann, R.S. (1998). Deep-sea epibenthic echinoderms and a temporally varying food supply: results from a one year time series in the N.E. Pacific. Deep-Sea Research II 45(1998): 817842.Google Scholar
Laurenson, C. H., Dobby, H., McLay, H. A. & Leslie, B. (2008). Biological features of the Lophius piscatorius catch in Scottish waters. ICES Journal of Marine Science 65: 12811290.Google Scholar
Laurenson, C.H., Hudson, I.R., Jones, D.O.B., & Priede, I.G. (2004). Deep water observations of anglerfish (Lophius piscatorius L.) in the North-eastern Atlantic Ocean by means of remotely operated vehicle. Journal of Fish Biology 65: 947960. doi: 10.1111/j.0022–1112.2004.00496.xCrossRefGoogle Scholar
Lawver, L.A. & Gahagan, L.M. (2003). Evolution of Cenozoic seaways in the circum-Antarctic region. Palaeogeography, Palaeoclimatology, Palaeoecology 198: 1137.CrossRefGoogle Scholar
Laxson, C.J., Condon, N.E., Drazen, J.C., & Yancey, P.H. (2011). Decreasing urea: trimethylamine N-oxide ratios with depth in chondrichthyes: a physiological depth limit? Physiological and Biochemical Zoology 84: 494505.CrossRefGoogle ScholarPubMed
Le Pichon, X. (1968). Sea-floor spreading and continental drift. Journal of Geophysical Research 73(12): 36613697.CrossRefGoogle Scholar
Lecointre, G., Gallut, C., Bonillo, C., Couloux, A., Ozouf-Costaz, C., & Dettaï, A. (2011). The Antarctic fish genus Artedidraco is paraphyletic (Teleostei, Notothenioidei, Artedidraconidae). Polar Biology 34(8): 11351145.Google Scholar
Levin, L.A. (2000). Oxygen minimum zone benthos: adaptation and community response to hypoxia. Oceanography and Marine Biology: An Annual Review 2003(41): 145.Google Scholar
Licht, M., Schmuecker, K., Huelsken, T., Hanel, R., Bartsch, P., & Paeckert, M. (2012). Contribution to the molecular phylogenetic analysis of extant holocephalan fishes (Holocephali, Chimaeriformes). Organisms Diversity and Evolution 12: 421432. doi: 10.1007/s13127-011-0071-1CrossRefGoogle Scholar
Lim, J., Fudge, D.S., Levy, N., & Gosline, J.M. (2006). Hagfish slime ecomechanics: testing the gill-clogging hypothesis. Journal of Experimental Biology 209: 702710.Google Scholar
Lindsay, D.J., Hunt, J.C., & Hayashi, K. (2001). Associations in the midwater zone: the penaeid shrimp Funchalia sagamiensis Fujino 1975 and pelagic tunicates (Order: Pyrosomatida). Marine and FreshWater Behaviour and Physiology 2001: 157170.CrossRefGoogle Scholar
Linley, T., Gerringer, M.E., Yancey, P.H., Drazen, J.C., Weinstock, C.L., & Jamieson, A.J. (2016). Fishes of the hadal zone including new species, in situ observations and depth records of Liparidae. Deep-sea Research I 114: 99110.CrossRefGoogle Scholar
Linley, T.D., Alt, C.H.S., Jones, D.O.B., & Priede, I.G. (2013). Bathyal demersal fishes of Charlie Gibbs Fracture Zone region (49°-54°N) of the Mid-Atlantic Ridge: III. Results from remotely operated vehicle (ROV) video transects. Deep-Sea Research II 98: 407411. http://dx.doi.org/10.1016/j.dsr2.2013.08.013CrossRefGoogle Scholar
Linley, T.D., Stewart, A.L., McMillan, P.J., Clark, M.R., Gerringer, M.E., Drazen, J.C., Fujii, T., & Jamieson, A.J. (2016). Bait attending fishes of the abyssal zone and hadal boundary: community structure, functional groups and species distribution in the Kermadec, New Hebrides and Mariana trenches. Deep-Sea Research I 121:3853.CrossRefGoogle Scholar
Linnaeus, C. (1735). Systemae Naturae. Leiden: Haak.Google Scholar
Linnaeus, C. (1758). Systemae Naturae, 10th edition. Stockholm: Salvius.Google Scholar
Lins, L.S.F., Ho, S.Y.W., Wilson, G.D.F., & Lo, N. (2012). Evidence for Permo-Triassic colonization of the deep sea by isopods. Biology Letters. doi:10.1098/rsbl.2012.0774CrossRefGoogle Scholar
Liu, J.Y. (2013). Status of marine biodiversity of the China Seas. PLoS ONE 8(1): e50719. http://doi.org/10.1371/journal.pone.0050719Google Scholar
Livermore, R., Nankivell, A., Eagles, G., & Morris, P. (2005) Paleogene opening of Drake Passage. Earth and Planetary Science Letters 236(2005): 459470.Google Scholar
Lloris, D., Matallanas, J., & Oliver, P. (2005). Hakes of the World (Family Merlucciidae): An Annotated and Illustrated Catalogue of Hake Species Known to Date. Rome: Food and Agriculture Organization of the United Nations. Available for download at www.fao.orgGoogle Scholar
Lloris, D., Stefanescu, C., & Rucabado, J. (1994). New data on the distribution and biology of Rhynchogadus hepaticus and Eretmophorus kleinenbergi (Osteichthyes: Moridae). Cybium 18: 129134.Google Scholar
Locket, N.A. (1977). Adaptations to the deep-sea environment. In: Crescitelli, F. (ed.) The Visual System in Vertebrates, pp. 67192. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Løkkeborg, S. (2005). Impacts of Trawling and Scallop Dredging on Benthic Habitats and Communities. FAO Fisheries Technical Paper. No. 472. Rome, FAO.Google Scholar
Lombarte, A. & Cruz, A. (2007). Otolith size trends in marine fish communities from different depth strata. Journal of Fish Biology 71: 5376.CrossRefGoogle Scholar
Longhurst, A. (1998). Ecological Geography of the Sea. New York: Academic Press.Google Scholar
Lopez, J.A., Westneat, M.W., & Hanel, R. (2007). The phylogenetic affinities of the mysterious anguilliform genera Coloconger and Thalassenchelys as supported by mtDNA sequences. Copeia 2007(4): 959966.CrossRefGoogle Scholar
Lorance, P., Large, P.A., Bergstad, O.A., & Gordon, J.D.M. (2008). Grenadiers of the NE Atlantic-distribution, biology, fisheries and the impacts, and developments in stock assessment and management. American Fisheries Society Symposium 63: 365397.Google Scholar
Love, R.H., Fisher, R.A., Wilson, M.A., & Nero, R.W. (2004). Unusual swimbladder behavior of fish in the Cariaco Trench. Deep-Sea Research I 51: 116.CrossRefGoogle Scholar
Lowe, R. T. (1833). Description of a new genus of acanthopterygian fishes, Alepisaurus ferox. Proceedings of the Zoological Society of London 1: 104.Google Scholar
Lowe, R.T. (1843–1860). A History of the Fishes of Madeira, London: Van Voorst.Google Scholar
Ludwig, H. & Macdonald, A.G. (2005). The significance of the activity of dissolved oxygen, and other gases, enhanced by high hydrostatic pressure. Comparative Biochemistry and Physiology 140A: 387395.CrossRefGoogle ScholarPubMed
Lundsten, L., Johnson, S.B., Cailliet, G.M., DeVogelaere, A.P,. & Clague, D.A. (2012). Morphological, molecular, and in situ behavioural observations of the rare deep-sea anglerfish Chaunacops coloratus (Garman, 1899), order Lophiiformes,in the eastern North Pacific. Deep-Sea Research I 68: 4653.CrossRefGoogle Scholar
Lütken, C. F. (1892). Spolia Atlantica. Scopelini Musei zoologici Universitatis Hauniensis. Bidrag til Kundskab om det aabne Havs Laxesild eller Scopeliner. Med et tillaeg om en anden pelagisk fiskeslaegt. Kongelige Danske Videnskabernes Selskab Series 6(3): 221297.Google Scholar
Machete, M., Morato, T., & Menezes, G. (2011). Experimental fisheries for black scabbardfish (Aphanopus carbo) in the Azores, Northeast Atlantic. ICES Journal of Marine Science 68: 302308.Google Scholar
Machida, Y (1989). Record of Abyssobrotula galatheae (Ophidiidae: Ophidiiformes) from the Izu-Bonin Trench, Japan. Bulletin of Marine Science and Fisheries Kochi University, Japan 11: 2325.Google Scholar
Machida, Y. & Hashimoto, J. (2002). Pyrolycus manusanus, a new genus and species of deep-sea eelpout from a hydrothermal vent field in the Manus Basin, Papua New Guinea (Zoarcidae, Lycodinae). Ichthyological Research 49:16.Google Scholar
Machida, Y. & Shiogaki, M. (1988). Leptochilichthys microlepis, a new species of the family Leptochilichthyidae, Salmoniformes, from Aomori, northern Japan. Japanese Journal of Ichthyology 35: 16.Google Scholar
Macpherson, E. (1985). Daily ration and feeding periodicity of some fishes off the coast of Namibia. Marine Ecology Progress Series 26: 253–60.CrossRefGoogle Scholar
Macpherson, E. (1989). Influence of geographical distribution, body size and diet on population density of benthic fishes off Namibia (South West Africa). Marine Ecology Progress Series 50: 295299.Google Scholar
Maguire, J.-J., Sissenwine, M., Csirke, J., Grainger, R., & Garcia, S. (2006). The State of World Highly Migratory, Straddling and Other High Seas Fishery Resources and Associated Species. FAO Fisheries Technical Paper. No. 495. Rome: FAO.Google Scholar
Mann, D.A. & Jarvis, S.M. (2004). Potential sound production by a deep-sea fish. Journal of the Acoustical Society of America 115: 23312333.Google Scholar
Mantyla, A.W. & Reid, J.L. (1983). Abyssal characteristics of the world ocean waters. Deep-Sea Research 30A: 805833.CrossRefGoogle Scholar
Marine Stewardship Council. (2015). New Zealand EEZ ling trawl and longline fishery. www.msc.org/track-a-fishery/fisheries-in-the-program/certified/pacific/new-zealand-eez-ling-trawl-and-longline-fishery accessed 15 September 2015.Google Scholar
Markle, D.F. & Merrett, N.R. (1980). The abyssal alepocephalid, Rinoctes nasutus (Pisces: Salmoniformes), a redescription and an evaluation of its systematic position. Journal of Zoology, London 190: 225239.Google Scholar
Markle, D.F. & Olney, J.E.(1990). Systematics of the pearlfishes (Pisces: Carapidae). Bulletin of Marine Science 47: 269410.Google Scholar
Markle, D.F & Sazonov, Y.I. (1996). Review of the rare deep-sea genus, Aulastomatomorpha (Teleostei: Salmoniformes), with a discussion of relationships. Copeia 1996(2): 497500.CrossRefGoogle Scholar
Marques, V., Chaves, C., Morai, A., Cardador, F., & Stratoudakis, Y. (2005). Distribution and abundance of snipefish (Macroramphosus spp.) off Portugal (1998–2003). Scientia Marina 69: 563576.Google Scholar
Marshall, A., Bennett, M.B., Kodja, G., Hinojosa-Alvarez, S., Galvan-Magana, F., et al. (2011). Manta birostris. IUCN 2011. IUCN Red List of Threatened Species. Version 2011.2. Available: <www.iucnredlist.org>. Accessed: 17 February 2015. www.iucnredlist.org/apps/redlist/details/198921/0..+Accessed:+17+February+2015.+www.iucnredlist.org/apps/redlist/details/198921/0.>Google Scholar
Marshall, N.B. (1953). Egg size in Arctic, Antarctic and deep-sea fishes. Evolution 7: 328341.CrossRefGoogle Scholar
Marshall, N.B. (1954). Aspects of Biology. London: Hutchinson.Google Scholar
Marshall, N.B. (1960). Swimbladder structure of deep-sea fishes in relation to their systematics and biology. Discovery Reports 31: 1122.Google Scholar
Marshall, N.B. (1965). Systematic and biological studies of the Macrourid fishes (Anacanthini-Teleostii) Deep-Sea Research 12: 299322.Google Scholar
Marshall, N.B. (1966a). Bathyprion danae a new genus and species of alepocephaliform fishes. Dana-Reports 68: 19.Google Scholar
Marshall, N.B. (1966b). The Life of Fishes. London: Weidenfeld and Nicolson.Google Scholar
Marshall, N. B. (1967). The olfactory organs of bathypelagic fishes. In: Aspects of Marine Zoology, pp. 5770. London: Academic Press.Google Scholar
Marshall, N.B. (1972). Swimbladder organisation and deep ranges of deep-sea teleosts. Symposia of the Society of Experimental Biology 26: 261272.Google Scholar
Marshall, N.J. (1996). The lateral line systems of three deep-sea fish. Journal of Fish Biology 49 (A): 239258.Google Scholar
Martín, J., Puig, P., Masqué, P., Palanques, A., & Sánchez-Gómez, A. (2014). Impact of bottom trawling on deep-sea sediment properties along the flanks of a submarine canyon. PLoS ONE 9(8): e104536. doi:10.1371/journal.pone.0104536CrossRefGoogle ScholarPubMed
Martini, F.H. & Beulig, A. (2013). Morphometics and gonadal development of the hagfish Eptatretus cirrhatus in New Zealand. PLoS ONE 8(11): e78740. doi:10.1371/journal.pone.0078740Google Scholar
Martins, R. & Fereira, C. (1995). Line fishing for black scabbardfish (Aphanopus carbo Lowe, 1839) and other deep water species in the eastern mid-Atlantic to the north of Madeira. In: Hooper, A.G. (ed.) Deep-Water Fisheries of the North Atlantic Oceanic Slope, pp. 323335. Dordrecht, Netherlands: Kluwer Academic Publishers.Google Scholar
Matallanas, J. (2000). On Mediterranean and some north-eastern Atlantic Liparidae (Pisces: Scorpaeniformes) with the restoration of Eutelichthys. Journal of the Marine Biological Association of the UK 80: 935939.Google Scholar
Matschiner, M., Colombo, M., Damerau, M., Ceballos, S., Hanel, R., & Salzburger, W. (2015). The adaptive radiation of Notothenioid fishes in the waters of Antarctica. In Riesch, R. et al. (eds.) Extremophile Fishes, pp. 3557. Heidelberg, Germany: Springer International. doi:10.1007/78-3-319-13362-1_3CrossRefGoogle Scholar
Matsui, T. & Rosenblatt, R.H. (1987). Review of the deep-sea fish family Platytroctidae (Pisces: Salmoniformes). Bulletin of the Scripps Institution of Oceanography 26: 1158. http://escholarship.org/uc/item/35v4k0ksGoogle Scholar
Matsunaga, T. & Rahman, A. (1998). What brought the adaptive immune system to vertebrates? The jaw hypothesis and the seahorse. Immunology Reviews 166: 177186.Google Scholar
Matsuura, K. (2015). Taxonomy and systematics of tetraodontiform fishes: a review focusing primarily on progress in the period from 1980 to 2014. Ichthyological Research 62(1): 72113.CrossRefGoogle Scholar
Mauchline, J. & Gordon, J.D.M. (1984a). Feeding and bathymetric distribution of the gadoid and morid fish of the Rockall Trough. Journal of the Marine Biological Association of the UK 64: 657665.Google Scholar
Mauchline, J. & Gordon, J.D.M. (1984b). Occurrence and feeding of berycomorphid and percomorphid teleost fish in the Rockall Trough. Journal du Conseil - Conseil international pour l’exploration de la mer 41: 239247.Google Scholar
Mauchline, J. & Gordon, J.D.M. (1991). Oceanic pelagic prey of benthopelagic fish in the benthic boundary layer of a marginal oceanic region. Marine Ecology Progress Series 74: 109115.Google Scholar
Maury, M.F. (1855). The Physical Geography of the Sea. New York: Harper and Bros.Google Scholar
McClain, C.R. & Hardy, S.M. (2010). The dynamics of biogeographic ranges in the deep sea. Proceedings of the Royal Society of London B 277: 35333546. doi:10.1098/rspb.2010.1057Google ScholarPubMed
McClain, C.R., Rex, M.A., & Jabbour, R. (2005). Deconstructing bathymetric patterns of body size in deep-sea gastropods. Marine Ecology Progress Series 297, 181187.CrossRefGoogle Scholar
McCleave, J.D. & Miller, M.J. (1994). Spawning of Conger oceanicus and Conger triporiceps (Congridae) in the Sargasso Sea and subsequent distribution of leptocephali. Environmental Biology of Fishes 39: 339355.CrossRefGoogle Scholar
McDowall, R.M. & Stewart, A.L. (1999). Further specimens of Agrostichthys parkeri (Teleostei: Regalecidae), with natural history notes. Proceedings of the 5th Indo-Pacific Fish Conference, Noumea, 1997, pp. 165174.Google Scholar
McEachran, J.D. & Dunn, K.A. (1998). Phylogenetic Analysis of Skates, a Morphologically Conservative Clade of Elasmobranchs (Chondrichthyes: Rajidae). Copeia 1998 (2): 271290.CrossRefGoogle Scholar
McIntyre, F.D., Collie, N., Stewart, M., Scala, L., & Fernandes, P.G. (2013). A visual survey technique for deep water fish: Estimating anglerfish abundance in closed areas. Journal of Fish Biology 83: 739753.CrossRefGoogle ScholarPubMed
McKenzie, A. (2016). Assessment of hoki (Macruronus novaezelandiae) in 2015. New Zealand Fisheries Assessment Report 2016/01. Wellington: Ministry for Primary Industries.Google Scholar
McMillan, P., Iwamoto, T., Stewart, A., & Smith, P.J. (2012). A new species of grenadier, genus Macrourus (Teleostei, Gadiformes, Macrouridae) from the southern hemisphere and a revision of the genus. Zootaxa 3165: 124.Google Scholar
McMillan, P.J., Griggs, L.H., Francis, M.P., Marriott, P.J, Paul, L.J., Mackay, E., Wood, B.A., Sui, H., & Wei, F. (2011). New Zealand Fishes. Volume 3: A Field Guide to Common Species Caught by Surface Fishing. New Zealand Aquatic Environment and Biodiversity Report No. 69. Wellington: Ministry of Fisheries.Google Scholar
McMillan, P.J. & Paulin, C.D. (1993). Descriptions of nine new species of rattails of the Genus Caelorinchus (Pisces, Macrouridae) from New Zealand. Copeia 1993: 819840.CrossRefGoogle Scholar
Mead, G.W., Bertelsen, E., & Cohen, D.M. (1964). Reproduction among deep-sea fishes. Deep-Sea Research 11: 569596.Google Scholar
Meléndez, R.C. & Markle, D.F. (1997). Phylogeny and zoogeography of Laemonema and Guttigadus (Pisces: Gadiformes: Moridae). Bulletin of Marine Science 61(3): 593670.Google Scholar
Melo, M.R.S. (2009a). Revision of the genus Chiasmodon (Acanthomorpha: Chiasmodontidae), with description of two new species. Copeia 2009(3): 583608.Google Scholar
Melo, M.R.S. (2009b). Taxonomic and phylogenetic revision of the family Chiasmodontidae (Perciformes: Acanthomorpha). PhD thesis Auburn University, Alabama, USA.Google Scholar
Merrett, N.R. (1981). First record of a Notacanth leptocephalus; a benthopelagic capture in the slope waters off Ireland. Journal of Fish Biology 18: 5357.Google Scholar
Merrett, N.R. (1987). A zone of faunal change in assemblages of abyssal demersal fish in the eastern North Atlantic: a response to seasonality in production? Biological Oceanography 5(2): 137151.Google Scholar
Merrett, N.R. (1994). Reproduction in the North Atlantic oceanic ichthyofauna and the relationship between fecundity and species’ sizes. Environmental Biology of Fishes 41: 207245.CrossRefGoogle Scholar
Merrett, N.R. & Barnes, S.H. (1996). Preliminary survey of egg envelope morphology in the Macrouridae and the possible implications of its ornamentation. Journal of Fish Biology 48: 101119.CrossRefGoogle Scholar
Merrett, N.R., Gordon, J.D.M., Stehman, M., & Haedrich, R.L. (1991a). Deep demeral fish assemblage structure in the Porcupine Seabight (Eastern North Atlantic): slope sampling by three different trawls compared. Journal of the Marine Biological Association of the UK 71: 329358.Google Scholar
Merrett, N.R. & Haedrich, R.L. (1997). Deep-Sea Demersal Fish & Fisheries. London: Chapman & Hall.Google Scholar
Merrett, N.R., Haedrich, R.L., Gordon, J.D.M., & Stehman, M. (1991b). Deep demeral fish assemblage structure in the Porcupine Seabight (Eastern North Atlantic): results of single warp trawling at lower slope to abyssal soundings. Journal of the Marine Biological Association of the UK 71: 359373.CrossRefGoogle Scholar
Merrett, N.R. & Marshall, N.B. (1981). Observations on the ecology of deep-sea bottom-living fishes collected off northwest Africa (08°–27°N). Progress in Oceanography 9: 185244.Google Scholar
Merrett, N.R. & Moore, J.A. (2005). A new genus and species of deep demersal fish (Teleostei: Stephanobercidae) from the tropical eastern North Atlantic. Journal of Fish Biology 67: 16991710.Google Scholar
Merrett, N.R. & Nielsen, J.G. (1987). A new genus and species of the family Ipnopidae (Pisces, Teleostei) from the eastern North Atlantic, with notes on its ecology. Journal of Fish Biology 31: 451464.CrossRefGoogle Scholar
Merrett, N.R., Sazonov, Y.I., & Shcherbachev, Y.N. (1983). A new genus and species of rattail fish (Macrouridae) from the eastern North Atlantic and the eastern Indian Ocean, with notes on its ecology. Journal of Fish Biology 22: 549561.CrossRefGoogle Scholar
Mienert, J., Berndt, C., Laberg, J.S., & Vorren, T.O. (2003). Slope instability of continental margins. In: Wefer, G., Billett, D., Hebbeln, D., Jørgensen, B.B., Schlüter, M., & van Weering, T. eds.) Ocean Margin Systems, pp. 179193. Berlin: Springer-Verlag.Google Scholar
Miller, M.J., Chikaraishi, Y., Ogawa, N.O., Yamada, Y., Tsukamoto, K., & Ohkouchi, N. (2013). A low trophic position of Japanese eel larvae indicates feeding on marine snow. Biology Letters 9: 20120826. http://dx.doi.org/10.1098/rsbl.2012.0826Google Scholar
Miller, M.J. & McCleave, J.D. (2007). Species assemblages of leptocephali in the southwestern Sargasso Sea. Marine Ecology Progress Series 344: 197212. doi: 10.3354/meps06923CrossRefGoogle Scholar
Miller, R.R. (1947). A new genus and species of deep-sea fish of the family Myctophidae from the Philippine Islands. Proceedings of the US National Museum 97: 8190.CrossRefGoogle Scholar
Milligan, R.J., Morris, K.J., Bett, B.J., Durden, J.M., Jones, D.O.B., Robert, K., Ruhl, H.A., & Bailey, D.M. (2016). High resolution study of the spatial distributions of abyssal fishes by autonomous underwater vehicle. Scientific Reports 6, Article number: 26095. doi:10.1038/srep26095CrossRefGoogle ScholarPubMed
Milliken, D.M. & Houde, E.D. (1984). A new species of Bregmacerotidae (pisces), Bregmaceros cantori, from the Western Atlantic Ocean. Bulletin of Marine Science 35: 1119.Google Scholar
Mills, E.L. (1983). Problems of deep-sea biology: An historical perspective. In: Rowe, G.T. (ed.) The Sea, pp. 179. Volume 8: Deep-Sea Biology. Cambridge, MA: Harvard University Press.Google Scholar
Mitson, R.B. & Knudsen, H.P. (2003). Causes and effects of underwater noise on fish abundance estimation. Aquatic Living Resources 16: 255263.CrossRefGoogle Scholar
Miya, M. (1994). Cyclothone kobayashii, a new gonostomatid fish (Teleostei: Stomiiiformes) from the Southern Ocean, with notes on its ecology. Copeia 1994(1): 191204.CrossRefGoogle Scholar
Miya, M. & Markle, D.F. (1993). Bajacalifornia aequatoris, new species of Alepocephalid fish (Pisces: Salmoniformes) from the central equatorial Pacific. Copeia 1993(3): 743747.CrossRefGoogle Scholar
Miya, M. & Nielsen, J.G. (1991). A new species of the deep-sea fish genus Parabrotula (Parabrotulidae) from Sagami Bay with notes on its biology. Japanese Journal of Ichthyology 38: 15.Google Scholar
Miya, M., Pietsch, T.W., Orr, J.W., Arnold, R.J., Satoh, T.P., Shedlock, A.M., Ho, H.-C., Shimazaki, M., Yabe, M., & Nishida, M. (2010). Evolutionary history of anglerfishes (Teleostei: Lophiiformes): a mitogenomic perspective. BMC Evolutionary Biology 2010 10: 58. www.biomedcentral.com/1471–2148/10/58CrossRefGoogle ScholarPubMed
Mizusawa, N., Takami, M., & Fukui, A. (2015). Redescription of the spookfish Dolichopteryx anascopa Brauer 1901 (Argentinoidei: Opisthoproctidae). Ichthyological Research 62: 236239.Google Scholar
Mok, H.-K. (1978). Scale-feeding in Tydemania navigatoris (Pisces: Triacanthodidae). Copeia 1978(2): 338340.CrossRefGoogle Scholar
Møller, P.R. & Jones, W.J. (2007). Eptatretus strickrotti n. sp. (Myxinidae): first hagfish captured from a hydrothermal vent. Biological Bulletin 212: 5566.CrossRefGoogle Scholar
Møller, P.R. & Stewart, A.L. (2006). Two new species of eelpouts (Teleostei, Zoarcidae) of the genus Seleniolycus from the Ross Dependency, Antarctica. Zootaxa 1376: 5356.Google Scholar
Montgomery, J. & Pankhurst, N. (1997). Sensory physiology. In: Randall, D.J. & Farrell, A.P. (eds.) Deep-Sea Fishes, pp. 325249. San Diego: Academic Press.CrossRefGoogle Scholar
Moore, J.A. (2002). Upside-down swimming behavior in a whipnose anglerfish (Teleostei: Ceratioidei: Gigantactinidae). Copeia 2002: 11441146.CrossRefGoogle Scholar
Mora, C., Tittensor, D. P., & Myers, R. A. (2008). The completeness of taxonomic inventories for describing the global diversity and distribution of marine fishes. Proceedings of the Royal Society B 275: 149155.Google Scholar
Morato, T., Machete, M., Kitchingham, A., Tempura, F., Lai, S., Menezes, G., Pitcher, T.J., & Santos, R.S. (2008). Abundance and distribution of seamounts in the Azores. Marine Ecology Progress Series 357: 1721.Google Scholar
Morato, T., Watson, R., Pitcher, T.J., & Pauly, D. (2006). Fishing down the deep. Fish and Fisheries 7: 2434.CrossRefGoogle Scholar
Morita, T. (1999). Molecular phylogenetic relationships of the deep-sea fish genus Coryphaenoides (Gadiformes: Macrouridae) based on mitochondrial DNA. Molecular Phylogenetics and Evolution 13: 447454.Google Scholar
Morita, T. (2008). Comparative sequence analysis of myosin heavy chain proteins from congeneric shallow- and deep-living rattail fish (genus Coryphaenoides). Journal of Experimental Biology 211: 13621367.Google Scholar
Morris, K.J., Bett, B.J., Durden, J.M., Huvenne, V.A.I., Milligan, R., Jones, D.O.B., McPhail, S., Bailey, D.M., Robert, K., Bailey, D.M., & Ruhl, H.A. (2014). A new method for ecological surveying of the abyss using autonomous underwater vehicle photography. Limnology and Oceanography Methods 12: 795809.CrossRefGoogle Scholar
Moser, H.G. (1996a). Anoplopomatidae: sablefish and skilfish. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 807809. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Moser, H.G. (1996b). Caristiidae: manefishes or veilfins. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 973975. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Moser, H.G. (1996c). Chauliodontidae: viperfishes. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 297299. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Moser, H.G. (ed.). (1996d). Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations. Atlas No 33. La Jolla, CA: South West Fisheries CentreGoogle Scholar
Moser, H.G. (1996e). Idiacanthidae: blackdragons. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 325327. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Moser, H.G. (1996f). Malacosteidae: loosejaws. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 321323. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Moser, H.G. (1996g). Myctophiformes: lanternfishes. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 383385. Atlas No 33La Jolla, CA: South West Fisheries Centre.Google Scholar
La Jolla, CA: South West Fisheries Centre.Google Scholar
Moser, H.G. & Ahlstrom, D.E.H. (1996). Bathylagidae: Blacksmelts and Smoothtongues. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 188207. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Moser, H.G. & Ahlstrom, E.H. (1974). Role of larval stages in systematic investigations of marine teleosts: the myctophidae, a case study. Fishery Bulletin 72: 391413.Google Scholar
Moser, H.G., Ahlstrom, E.H., & Paxton, J.R. (1984). Myctophidae: development. In: Moser, H.G., Richards, W.J., Cohen, D.M., Fahay, M.P., Kendall, M.P., & Richardson, S.L. (eds.) Ontogeny and Systematics of Fishes, pp. 218239. Special Publication. Lawrence, KS: American Society of Ichthyologists and Herpetologists.Google Scholar
Moser, H.G. & Charter, S.R. (1996). Notacanthiformes. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 8285. Atlas No 33. La Jolla, CA: South West Fisheries Centre, La Jolla.Google Scholar
MSC. (2016). Marine Stewardship Council, Certified sustainable seafood. Online publication accessed 1 April 2016: www.msc.org/Google Scholar
Mueter, F.J., Reist, J.D., Majewski, A.R., Sawatzky, C.D., Christiansen, J.S., Hedges, K.J., Coad, B.W., Karamushko, O.V., Lauth, R.R., Lynghammar, A., MacPhee, S.A., & Mecklenburg, C.W. (2013). Marine fishes of the Arctic. Online publication www.arctic.noaa.gov/report13/marine_fish.html accessed 29 September 2016.Google Scholar
Müller, R. D., Sdrolias, M., Gaina, C., & Roest, W. R. (2008). Age, spreading rates, and spreading asymmetry of the world’s ocean crust. Geochemistry, Geophysics, Geosystems 9, Q04006. doi:10.1029/2007GC001743.Google Scholar
Munk, O., (1959). The eyes of Ipnops murrayi Günther, 1878. Galathea Reports 3: 7987.Google Scholar
Munk, O. (1965). Ocular degeneration in deep-sea fishes. Galathea Reports 8: 2131.Google Scholar
Munk, O. (1977). The visual cells and retinal tapetum of the foveate deep-sea fish Scopelosaurus lepidus (Teleostei). Zoomorphologie 87: 2149.CrossRefGoogle Scholar
Munk, O. & Frederiksen, R.D. (1974). On the function of aphakic apertures in teleosts. Videnskabelige Meddelelser Dansk Naturhistorisk Forening 137: 6594.Google Scholar
Munk, O. & Jørgensen, J.M. (2010). Putatively luminous tissue in the abdominal pouch of a male dalatiine shark, Euprotomicroides zantedeschia Hulley & Penrith, 1966. Acta Zoologica 69: 247251. doi: 10.1111/j.1463–6395.1988.tb00921.Google Scholar
Munroe, T.A. (1998). Systematics and ecology of tonguefishes of the genus Symphurus (Cynoglossidae: Pleuronectiformes) from the western Atlantic Ocean. Fishery Bulletin 96(1): 1182.Google Scholar
Munroe, T.A. & Hashimoto, J. (2008). A new Western Pacific tonguefish (Pleuronectiformes: Cynoglossidae): The first Pleuronectiform discovered at active hydrothermal vents. Zootaxa 1839: 4359.Google Scholar
Munroe, T.A., Tyler, J., & Tunnicliffe, V. (2011). Description and biological observations on a new species of deep water symphurine tonguefish (Pleuronectiformes: Cynoglossidae: Symphurus) collected at Volcano–19, Tonga Arc, West Pacific Ocean. Zootaxa 3061: 5366.CrossRefGoogle Scholar
Murray, J. & Hjort, J. (1912). The Depths of the Ocean. London: Macmillan.Google Scholar
Murray, J.W., Jannasch, H.W., Honjo, S., Anderson, R.F., Reeburgh, W.S., Top, Z., Friederich, G.E., Codispoti, L.A., & Izdar, E. (1989). Unexpected changes in the oxic/anoxic interface in the Black Sea. Nature 338: 411413. doi:10.1038/338411a0CrossRefGoogle Scholar
Musick, J.A., Bruton, M.N., & Balon, E. K. (eds.). (1991). The biology of Latimeria chalumnae and evolution of coelocanths. Environmental Biology of Fishes 32(1–4): 1445.Google Scholar
Musick, J. A. & Cotton, C. F. (2015). Bathymetric limits of chondrichthyans in the deep sea: A re-evaluation. Deep-Sea Research II 115: 7380.CrossRefGoogle Scholar
Mytilineou, C., Politou, C.-Y., Papaconstantinou, C., Kavadas, S., D’Onghia, G., & Sion, L. (2005). Deep-water fish fauna in the Eastern Ionian Sea. Belgian Journal of Zoology 135: 229233.Google Scholar
Nafpaktitis, B.G., Robertson, D.A., & Paxton, J.A. (1995). Four new species of Diaphus (Myctophidae) from the Indo-Pacific. New Zealand Journal of Marine and Freshwater Research 29: 335344.Google Scholar
Nakae, M. & Sasaki, K. (2002). A scale-eating triacanthodid, Macrorhamphosodes uradoi: prey fishes and mouth ‘handedness’ (Tetraodontiformes, Triacanthoidei). Ichthyological Research 49: 714.CrossRefGoogle Scholar
Nakamura, I., Meyer, C.G. & Sato, K. (2015). Unexpected positive buoyancy in deep sea sharks, Hexanchus griseus, and an Echinorhinus cookei. PLoS ONE 10(6): e0127667. doi:10.1371/journal.pone.0127667Google Scholar
Nakamura, I. & Parin, N.V. (1993). FAO Species Catalogue. Vol. 15. Snake mackerels and cutlassfishes of the world (families Gempylidae and Trichiuridae). An annotated and illustrated catalogue of the snake mackerels, snoeks, escolars, gemfishes, sackfishes, domine, oilfish, cutlassfishes,. scabbardfishes, hairtails, and frostfishes known to date. FAO Fisheries Synopsis 125(15): 136.Google Scholar
Nakanishi, M. & Hashimoto, J. (2011). A precise bathymetric map of the world’s deepest seafloor, Challenger Deep in the Mariana Trench. Marine Geophysical Researches 32(4): 455463. doi:10.1007/s11001-011–9134-0Google Scholar
Nashida, K., Sakaji, H., & Honda, H. (2007). Spawning seasons of adult and growth of 0-year-old Deepsea smelt Glossanodon semifasciatus in Tosa Bay, Pacific coast of Shikoku. Bulletin of the Japanese Society of Fisheries and Oceanography 71(4): 270278.Google Scholar
National Geographic. (1967). Indian Ocean Floor Map. October 1967. Washington DC.Google Scholar
National Geographic. (1968). Atlantic Ocean Floor Map. June 1968. Washington DC.Google Scholar
National Geographic. (1969). Pacific Ocean Floor Map. October 1969. Washington DC.Google Scholar
National Marine Fisheries Service. (2013). U.S. National Bycatch Report First edition Update 1. In: Benaka, L.R., Rilling, C., Seney, E.E., & Winarsoo, H. (eds.). Washington, DC: U.S. Department of Commerce. www.st.nmfs.noaa.gov/observer-home/first-edition-update-1 accessed 16 December 2015.Google Scholar
Near, T.J, Dornburg, A., Eytan, R.I., Keck, B.P., Smith, W.L., Kuhn, K.L., Moore, J.A. Price, A.A., Burbrink, F.T, Friedman, M., & Wainwright, P.C. (2013). Phylogeny and tempo of diversification in the superradiation of spiny-rayed fishes. Proceedings of the National Academy of Sciences of the USA 1101: 273812743. doi:10.1073/pnas.1304661110Google Scholar
Near, T. J., Eytan, R. I., Dornburg, A., Kuhn, K. L., Moore, J. A., Davis, M. P., Wainwright, P. C., Friedman, M., & Smith, W. L. (2012). Resolution of ray-finned fish phylogeny and timing of diversification. Proceedings of the National Academy of Sciences of the United States of America 109: 1369813703. doi: 10.1073/pnas.1206625109CrossRefGoogle ScholarPubMed
Near, T.J., Jones, C.D., & Eastman, J.T. (2009). Geographic intraspecific variation in buoyancy within Antarctic notothenioid fishes. Antarctic Science 21(2): 123129.Google Scholar
Neat, F. & Burns, F. (2010). Stable abundance, but changing size structure in grenadier fishes (Macrouridae) over a decade (1998–2008) in which deepwater fisheries became regulated. Deep Sea Research Part I Oceanographic Research Papers 57: 434440. doi: 10.1016/j.dsr.2009.12.003CrossRefGoogle Scholar
Neat, F.C. & Campbell, N. (2013). Proliferation of elongate fishes in the deep sea. Journal of Fish Biology 83: 15761591.CrossRefGoogle ScholarPubMed
Neira, F.J. & Lyle, J.M. (2011). DEPM-based spawning biomass of Emmelichthys nitidus (Emmelichthyidae) to underpin a developing mid-water trawl fishery in south-eastern Australia. Fisheries Research 110: 236243.Google Scholar
Nelson, J. S. (1994). Fishes of the World, 3rd edition. New York: Wiley.Google Scholar
Nelson, J.S. (2006). Fishes of the World, 4th edition. Hoboken, NJ: Wiley.Google Scholar
Nemeth, D. (1994). Systematics and distribution of fishes of the family Champsodontidae (Teleostei: Perciformes), with descriptions of three new species. Copeia 1994(2): 347371.Google Scholar
New Zealand. (2011). Southern Blue Whiting, Fisheries Chapter Plan. Online publication. www.mpi.govt.nz. Accessed 18 August 2015.Google Scholar
Niass, F. & Ozawa, T. (2000). Morphological differences between North Pacific and Atlantic specimens of Lampadena anomala (Family Myctophidae). Ichthyological Research 47: 299302.Google Scholar
Nicol, J.A.C. (1958). Observations on luminescence in pelagic animals. Journal of the Marine Biological Association of the UK 37(3): 705752.Google Scholar
Nicol, J.A.C. (1960). Studies on luminescence. On the subocular light-organs of stomiatoid fishes. Journal of the Marine Biological Association of the UK 39: 529548.Google Scholar
Nielsen, J., Badcock, J., & Merrett, N.R. (1990). New data elucidating the taxonomy and ecology of the Parabrotulidae (Pisces: Zoarcoidei). Journal of Fish Biology 37: 437448.CrossRefGoogle Scholar
Nielsen, J., Hedeholm, R.B., Heinemeier, J., Bushnell, P.G., Christiansen, J.S., Olsen, J., Ramsey, C.W., Brill, R.W., Simon, M., Steffensen, K.F., & Steffensen, J.F. (2016). Eye lens radiocarbon reveals centuries of longevity in the Greenland shark (Somniosus microcephalus). Science 353: 702704. doi: 10.1126/science.aaf1703CrossRefGoogle ScholarPubMed
Nielsen, J.G. (1964). Fishes from depths exceeding 6000 meters. Galathea Report 7: 113124.Google Scholar
Nielsen, J. G. (1966). Synopsis of the Ipnopidae (Pisces, Iniomi) with description of two new abyssal species. Galathea Report 8: 4975.Google Scholar
Nielsen, J. G. (1977). The deepest living fish Abyssobrotula galathea. A new genus and species of oviparous ophidiids (Pisces, Brotulidae). Galathea Rep. 14: 4148.Google Scholar
Nielsen, J.G. & Cohen, D.M. (2005). Thermichthys (Bythitidae), a replacement name for preoccupied Gerhardia Nielsen & Cohen, 2002 and a second specimen of Thermichthys hollisi from the southeast Pacific. Cybium 29(4): 395398.Google Scholar
Nielsen, J.G., Cohen, D.M., Markle, D.F. & Robins, C.R. (1999). FAO species catalogue. Volume 18. Ophidiiform fishes of the world (Order Ophidiiformes). An annotated and illustrated catalogue of pearlfishes, cusk-eels, brotulas and other ophidiiform fishes known to date. FAO Fisheries Synopsis. No. 125, Vol. 18. Rome: FAO.Google Scholar
Nielsen, J.G. & Hartel, K.E. (1996). Monognathus berteli sp. nov. from the Indian Ocean (Pisces, Monognathidae). Ichthyological Research 43(2): 113115.CrossRefGoogle Scholar
Nielsen, J.G. & Larsen, V. (1968). Synopsis of the Bathylaconidae (Pisces, Isospondyli) with a new eastern Pacific species. Galathea Reports 9: 221238.Google Scholar
Nielsen, J.G. & Machida, Y. (2006). Neobythitoides serratus, a new bathyal genus and species from the East China Sea (Teleostei: Ophidiidae). Zootaxa 1227: 6368.Google Scholar
Nielsen, J.G. & Merrett, N.R. (2000). Revision of the cosmopolitan deep-sea Genus Bassozetus (Pisces: Ophiliiedae) with two new species. Galathea Report 18: 756.Google Scholar
Nielsen, J.G., Mincarone, M.M., & Di Dario, F. (2015). A new deep-sea species of Barathronus Goode & Bean from Brazil, with notes on Barathronus bicolor Goode & Bean (Ophidiiformes: Aphyonidae). Neotropical. Ichthyology 13: 5360.Google Scholar
Nielsen, J.G. & Møller, P.R. (2011). Revision of the bathyal cusk-eels of the genus Bassogigas (Ophidiidae) with description of a new species from off Guam, west Pacific Ocean. Journal of Fish Biology 78 : 783795.CrossRefGoogle ScholarPubMed
Nielsen, J.G., Møller, P.R., & Segonzac, M. (2006). Ventichthys biospeedoi n. gen. et sp. (Teleostei, Ophidiidae) from a hydrothermal vent in the South East Pacific. Zootaxa 1247: 1324.CrossRefGoogle Scholar
Nielsen, J.G., Ross, S.W. & Cohen, D.M. (2009). Atlantic occurrence of the genus Bellottia (Teleostei, Bythitidae) with two new species from the Western North Atlantic. Zootaxa 2018: 4557.Google Scholar
Nikaido, M., Sasaki, T., Emerson, J.J., Aibara, M., Mzighani, S.I., Budeba, Y.L., Ngatunga, B.P., Iwata, M., Abe, Y., Li, W.H., & Okada, N. (2011). Genetically distinct coelacanth population off the northern Tanzanian coast. Proceedings of the National Academy of Sciences of the USA 108: 1800918013. doi: 10.1073/pnas.1115675108Google Scholar
Nishimura, S. (1966). Early life history of the deep-sea smelt, Glossanodon semifasciatus (Kishinouye) (Teleostei: Clupeida) PART I. Publications of the Seto Marine Biological Laboratory XIII(5): 349360.Google Scholar
Norse, E.A., Brooke, S., Cheung, W.W.L., Clark, M.R., Ekelandd, I., Froese, R., Gjerde, K.M., Haedrich, R.L., Heppell, S.S., Morato, T., Morgan, L.E., Pauly, D., Sumaila, R., & Watson, R. (2012). Sustainability of deep-sea fisheries. Marine Policy 36: 307320.CrossRefGoogle Scholar
Nunoo, F., Bannermann, P., Russell, B., & Poss, S. (2015). Helicolenus dactylopterus. The IUCN Red List of Threatened Species 2015: e.T195093A15592445. http://dx.doi.org/10.2305/IUCN.UK.2015-4.RLTS.T195093A15592445.en. Downloaded on 21 March 2017.CrossRefGoogle Scholar
Nybelin, O. (1957). Deep-sea bottom fishes. Report of the Swedish Deep Sea Expedition 2. Zoology 20: 247345.Google Scholar
NZ. (2009). History of fishing in New Zealand: growth and the EEZ. NZ Fisheries Info site: http://fs.fish.govt.nz/Page.aspx?pk=51&tk=166 accessed 6 May 2016.Google Scholar
O’Connor, S. & Veth, P. (2005). Early Holocene shell fish hooks from Lene Hara Cave, East Timor establish complex fishing technology was in use in Island South East Asia five thousand years before Austronesian settlement. Antiquity 79: 249256.Google Scholar
Okamura, O. (1970). Studies on the macrourid fishes of Japan. Morphology, ecology and phylogeny. Reports of the USA Marine Biological Station Kochi University. 17: 1179.Google Scholar
Okiyama, M., Tominaga, Y., & Ida, H. (2007). A megapterygium larva of Discoverichthys praecox (Aulopiformes: Ipnopidae) from the tropical western Pacific. Ichthyological Research 54: 262267. doi: 10.1007/s10228-007–0399-xCrossRefGoogle Scholar
O’Leary, B.C., Brown, R.L., Johnson, D.E., vonNordheim, H., Ardron, J., Packeiser, T., & Roberts, C.M. (2012). The first network of marine protected areas (MPAs) in the high seas: The process, the challenges and where next. Marine Policy 36: 598605.CrossRefGoogle Scholar
Olivar, M.P., Bernal, A., Molí, B., Peña, M., Balbín, R., Castellón, A., Miquel, J., & Massuti, E. (2012). Vertical distribution, diversity and assemblages of mesopelagic fishes in the western Mediterranean. Deep-Sea Research I 62: 5369.CrossRefGoogle Scholar
Olivar, M.P., González-Gordillo, J.I., Salat, J., Chust, G., Cózar, A., Hernández-León, A., Fernández de Puelles, M.L., & Irigoien, X. (2016). The contribution of migratory mesopelagic fishes to neuston fish assemblages across the Atlantic, Indian and Pacific Oceans. Marine and Freshwater Research 67: 11141127.CrossRefGoogle Scholar
Olney, J.E. (2005). Stylephoridae: Tube eyes. In: Richards, W.J. (ed.) Early Stages of Atlantic Fishes: An Identification Guide for the West Central North Atlantic. Volume 1, pp.10131014. Boca Raton, FL: CRC Press.Google Scholar
Olney, J.E., Johnson, G.D., & Baldwin, C.C. (1993). Phylogeny of lampridiform fishes. Bulletin of Marine Science 52: 137169.Google Scholar
Olson, K. R. (1996). Secondary circulation in fish: Anatomical organization and physiological significance. Journal of Experimental Zoology 275: 172185.3.0.CO;2-A>CrossRefGoogle Scholar
Oppo, D.W. & Curry, W.B. (2012). Deep Atlantic circulation during the last glacial maximum and deglaciation. Nature Education Knowledge 3(10): 1. www.nature.com/scitable/knowledge/library/deep-atlantic-circulation-during-the-last-glacial-25858002Google Scholar
Orlov, A.M. & Tokranov, A.M. (2011). Some rare and insufficiently studied snailfish (Liparidae, Scorpaeniformes, Pisces) in the PacificWaters off the Northern Kuril Islands and Southeastern Kamchatka, Russia. International Scholarly Research Network ISRN Zoology 2011, Article ID 341640, 12 pages doi:10.5402/2011/341640Google Scholar
Orr, J. W. & Busby, M.S. (2001). Prognatholiparis ptychomandibularis, a new genus and species of the fish family Liparidae (Teleostei: Scorpaeniformes) from the Aleutian Islands, Alaska. Proceedings of the Biological Society of Washington 114(1): 5157.Google Scholar
Overdick, A.A., Busby, M.S., & Blood, D.M. (2014). Descriptions of eggs of snailfishes (family Liparidae) from the Bering Sea and eastern North Pacific Ocean. Ichthyological Research 61(2): 131141. doi: 10.1007/s10228-013–0384-5Google Scholar
Owen, P. & Rice, T. (1999). Decommissioning of Brent Spar. London: Spon Press.CrossRefGoogle Scholar
Ozaka, C., Yamamoto, N., & Somiya, H. (2009). The Aglomerular Kidney of the Deep-sea Fish, Ateleopus japonicus (Ateleopodiformes: Ateleopodidae): Evidence of wider occurrence of the aglomerular condition in Teleostei. Copeia 2009: 609617.Google Scholar
Padilla, A., Zeller, D., & Pauly, D. (2015). The fish and fisheries of Bouvet Island. In: Palomares, M.L.D. & Pauly, D. (eds.) Marine Fisheries Catches of SubAntarctic Islands, 1950–2010, pp. 2130. Fisheries Centre Research Report 23(1). Vancouver, BC: Fisheries Centre, University of British Columbia.Google Scholar
Papaconstantinou, C., Anastasopoulou, K., & Caragitsou, E. (1997). Comments on the mesopelagic fauna of the North Aegean Sea. Cybium 21: 281288.Google Scholar
Parin, N.V. (2004). A new mesopelagic fish Ioichthys kashkini Parin, gen.et sp, nova (Opisthoproctidae) from the North Western part of the Indian Ocean. Journal of Ichthyology 44(7): 485488.Google Scholar
Parin, N.V., Belyanina, T.N., & Evseenko, S.A. (2009) Materials to the revision of the genus Dolichopteryx and closely related taxa (Ioichthys, Bathylychnops) with the separation of a new genus Dolichopteroides and Description of three new species (Fam. Opisthoproctidae). Journal of Ichthyology 49(10): 839851.CrossRefGoogle Scholar
Parin, N.V. & Borodulina, O.D. (2006). Antigonias (Antigonia, Caproidae) of the Western Indian Ocean: 2. Species with eight spiny rays in the dorsal fin. Journal of Ichthyology 46 3): 203211.Google Scholar
Parin, N.V. & Kobyliansky, S.G. (1996). Diagnoses and distribution of fifteen species recognized in genus Maurolicus Cocco (Sternoptychidae, Stomiiformes) with a key to their identification. Cybium 20(2): 185195.Google Scholar
Parmentier, E. & Vandewalle, P. (2003). Morphological adaptations of Pearlfish (Carapidae) to their various habitats. In: Val, A. L. & Kapoor, B. G. (eds.) Fish Adaptations, pp. 261276. Oxford: Science Publisher.Google Scholar
Parr, A.E. (1948). The classification of the fishes of the genera Bathylaco and Macromastax, possible intermediates between the Isospondyli and Iniomi. Copeia 1948: 4854.Google Scholar
Parrish, R.H. (1972). Symbiosis in blacktail snailfish, Careproctus melanurus, and box crab, Lopholithodes foraminatus. California Fish and Game 58: 239240.Google Scholar
Partridge, J.C., Douglas, R.H., Marshall, N.J., Chung, W.-S., Jordan, T.M., & Wagner, H.-J. (2014). Reflecting optics in the diverticular eye of a deep-sea barreleye fish (Rhynchohyalus natalensis). Proceedings of the Royal Society. B 281: 20133223. http://dx.doi.org/10.1098/rspb.2013.3223Google ScholarPubMed
Passow, U. & Carlson, C.A. (2012). The biological pump in a high CO2 world. Marine Ecology Progress Series 470: 249271.Google Scholar
Paulin, C.D. & Moreland, J.M. (1979). Congiopodus coriaceus, a new species of pigfish, and a redescription of C. leucopaecilus(Richardson), from New Zealand. (Pisces:Congiopodidae). New Zealand Journal of Zoology 6: 601608.Google Scholar
Pauly, D. (1995). Anecdotes and the shifting baseline syndrome of fisheries. Trends in Ecology and Evolution 10: 430.CrossRefGoogle ScholarPubMed
Pauly, D., Christensen, V., Dalsgaard, J., Froese, R., & Torres, F. Jr. (1998). Fishing down marine food webs. Science 279: 860863. doi: 10.1126/science.279.5352.860CrossRefGoogle ScholarPubMed
Pavlov, D.S., Parin, N.V., & Balushkin, A.V. (2009). In memory of Anatole Petrovich Andriashev (August 19, 1910 to January 4, 2009). Journal of Ichthyology 49: 547562. doi: 10.1134/S003294520907008XGoogle Scholar
Paxton, J.R. (1972). Osteology and relationships of the lanternfishes (family Myctophidae). Bulletin of the Natural History Museum of Los Angeles City 13: 181.Google Scholar
Paxton, J.R. (1989). Synopsis of the Whalefishes (Family Cetomimidae) with descriptions of four new genera. Records of the Australian Museum 41: 135206. ISSN 0067 1975CrossRefGoogle Scholar
Payne, A.I.L. & Punt, A.E. (1995). Biology and fisheries of South African Cape Hakes (M. capensis and M. paradoxus). In: Alheit, J. & Pitcher, T.J. (eds.) Hake, Biology, Fisheries and Markets, pp. 1547. London: Chapman & Hall.Google Scholar
Pearcy, W.G. & Ambler, J.W. (1974). Food habits of deep-sea macrourid fishes off the Oregon coast. Deep-Sea Research 21: 745759.Google Scholar
Pearcy, W.G., & Hubbard, L. (1964). A modification of the Isaacs–Kidd midwater trawl for sampling at different depth intervals. Deep-Sea Research 11: 263264.Google Scholar
Pearcy, W.G., Meyer, S.L., & Munk, O. (1965). A ‘four-eyed’ fish from the deep-sea: Bathylychnops exilis Cohen, 1958. Nature 207: 12601262.CrossRefGoogle ScholarPubMed
Pearcy, W.G., Stein, D.L., & Carney, R.S. (1982). The deep-sea benthic fish fauna of the Northeastern Pacific Ocean on Cascadia and Tufts Abyssal Plains and adjoining continental slopes. Biological Oceanography 1(4): 375428.Google Scholar
Pelster, B. (1997). Buoyancy at depth. In: Randall, D.J. & Farrell, A.P. (eds.) Deep-Sea Fishes, pp. 195237. Academic Press, San Diego, California.CrossRefGoogle Scholar
Pepperell, J. (2010). Fishes of the Open Ocean: A Natural History and Illustrated Guide. Chicago, IL: University of Chicago.Google Scholar
Pequeño, G. (2008). Grenadier fishes from Chilean waters: some aspects in relation to fisheries. American Fisheries Society Symposium 63: 4148.Google Scholar
Pérès, J.M. (1965). Apercu sur les résultats de deux plongées effectuées dans le ravin de Puerto-Rico par le bathyscaphe Archimède. Deep-Sea Research 12: 883891.Google Scholar
Petrov, A.F. (2011). New data on the diet of deepsea icefish Chionobathyscus dewitti (Channichthyidae) in the Ross Sea in 2010. Journal of Ichthyology 51(8): 692694.Google Scholar
Pettersson, O. & Drechsel, C.F. (1913). Memorandum on investigations in the Atlantic Ocean and programme for same. Rapports Et Procès – Verbaux Des Réunions Du Conseil Permanent International Pour L’exploration De La Mer 16: 121, 2 plates.Google Scholar
Pham, C.K., Canha, A., Diogo, H., Pereira, J.G., Prieto, R., & Morato, T. (2013). Total marine fishery catch for the Azores (1950–2010). ICES Journal of Marine Science, doi.10.1093/icesjms/fst024Google Scholar
Pham, C.K., Diogo, H., Menezes, G., Porteiro, F., Braga-Henriques, A., Vandeperre, F., & Morato, T. (2014). Deep-water longline fishing has reduced impact on Vulnerable Marine Ecosystems. Scientific Reports 4, 4837. doi:10.1038/srep04837 (2014)CrossRefGoogle ScholarPubMed
Phleger, C.E. & Grigor, M.R. (1990). Role of wax esters in determining buoyancy in Hoplostethus atlanticus (Beryciformes: Trachichthyidae). Marine Biology 105: 229233.Google Scholar
Pianka, E.R. (1970). On r and K selection. American Naturalist 104(940): 592597. doi:10.1086/282697CrossRefGoogle Scholar
Piccard, J. & Dietz, R.S. (1961). Seven Miles Down. London: Longmans.Google Scholar
Pietsch, T.W. (1978). The feeding mechanism of Stylephorus chordatus (Teleostei: Lampridiformes): functional and ecological implications. Copeia 1978(2): 255262.CrossRefGoogle Scholar
Pietsch, T.W. (2009). Oceanic Anglerfishes, Extraordinary Diversity in the Deep Sea. Berkeley: University of California Press.CrossRefGoogle Scholar
Pietsch, T.W., Ho, H.-C., & Chen, H.-M. (2004). Revision of the deep-sea anglerfish genus Bufoceratias Whitley (Lophiiformes: Ceratioidei: Diceratiidae), with description of a new species from the Indo-West Pacific Ocean. Copeia 2004(1): 98107.CrossRefGoogle Scholar
Pietsch, T.W. & Kenaley, C.P. (2011). A new species of deep-sea ceratioid anglerfish, genus Himantolophus (Lophiiformes: Himantolophidae), from southern waters of all three major oceans of the world. Copeia 2011(4): 490496.CrossRefGoogle Scholar
Pietsch, T.W. & Orr, J.W. (2007). Phylogenetic relationships of deep-sea anglerfishes of the suborder Ceratioidei (Teleostei: Lophiiformes) based on morphology. Copeia, 2007(1): 134.Google Scholar
Pietsch, T.W. & Shimazaki, M. (2005). Revision of the deep-sea anglerfish genus Acentrophyrne Regan (Lophiiformes: Ceratioidei: Linophrynidae), with the description of a new species from off Peru. Copeia 2005(2): 246251.CrossRefGoogle Scholar
Pietsch, T.W. & Van Duzer, J.P. (1980). Systematics and distribution of ceratioid anglerfishes of the family Melanocetidae with the description of a new species from the Eastern North Pacific Ocean. Fishery Bulletin 78: 5987.Google Scholar
Pinto, G. (2007) Global distribution of passive margins. Online Map www.unalmed.edu.co/rrodriguez/Passive%20Margin/Passive%20margin.htm accessed 23 September 2014.Google Scholar
Pirrera, L., Bottari, T., Busalacchi, B., Giordano, D., Modica, L., Perdichizzi, A., Perdichizzi, F., Profeta, A., & Rinelli, P. (2009). Distribution and population structure of the fish Helicolenus dactylopterus dactylopterus (Delaroche, 1809) in the Central Mediterranean (Southern Tyrrhenian Sea). Marine Ecology 30 (Suppl.1): 161174.CrossRefGoogle Scholar
Pollard, D. (2014). Microichthys coccoi. The IUCN Red List of Threatened Species 2014: e.T194859A49087606. http://dx.doi.org/10.2305/IUCN.UK.2014–3.RLTS.T194859A49087606.en. Downloaded on 24 November 2015.Google Scholar
Poltev Yu, N. (2013). Carcinophyly of fish of the genus Careproctus (Scorpaeniformes: Liparidae) in waters of Southeastern Sakhalin (Sea of Okhotsk). Journal of Ichthyology 53: 416424.Google Scholar
Pond, D.W., Fallick, A.E., Stevens, C.J., Morrison, D.J. & Dixon, D.R. (2008). Vertebrate nutrition in a deep-sea hydrothermal vent ecosystem: Fatty acid and stable isotope evidence. Deep-Sea Research I 55:17181726.Google Scholar
Pope, E.C., Hays, G.C., Thys, T.M., Doyle, T.K., Sims, D.W., Queiroz, N., Hobson, V.J., Kubicek, L., & Houghton, J.D.R. (2010). The biology and ecology of the ocean sunfish Mola mola: a review of current knowledge and future research perspectives Reviews in Fish Biology and Fisheries 20(4): 471487.CrossRefGoogle Scholar
Popper, A. N. & Fay, R. R. (1993). Sound detection and processing by fish: critical review and major research questions. Brain, Behaviour and Evolution 41: 1438.CrossRefGoogle ScholarPubMed
Porcu, C., Follesa, M.C., Gastoni, A., Mulas, A., Pedoni, C., & Cau, A. (2013). The reproductive cycle of a deep-sea eel, Nettastoma melanurum (Nettastomatidae:Anguilliformes) from the south-eastern Sardinian Sea (central-western Mediterranean). Journal of the Marine Biological Association of the UK 93(4): 11051115. doi:10.1017/S0025315412001452CrossRefGoogle Scholar
Post, A. & Quèro, J.-C. (1991.) Distribution et taxinomie de Howella (Perciformes, Perchthyidae) de L’Atlantique. Cybium 15(2): 111128.Google Scholar
Potthoff, T., Richards, W.J., & Ueyanagi, S. (1980). Development of Scombrolabrax heterolepis (Pisces: Scombrolabracidae) and comments on familial relationships. Bulletin Marine Science 30(2): 329357.Google Scholar
Poulsen, J.Y., Byrkjedal, I., Willassen, E., Rees, D., Takeshima, H., Satoh, T.P., Shinohara, G., Nishida, M., & Miya, M. (2013). Mitogenomic sequences and evidence from unique gene rearrangements corroborate evolutionary relationships of myctophiformes (Neoteleostei). BMC Evolutionary Biology 2013 13: 111. doi:10.1186/1471–2148-13-111Google ScholarPubMed
Powell, M.L., Kavanaugh, S.I., & Sower, S.A. (2005). Current knowledge of hagfish reproduction: implications for fisheries management. Integrative and Comparative Biology 45: 158165.CrossRefGoogle ScholarPubMed
Powell, S.M., Haedrich, R.L., & McEachran, J.D. (2003). The deep-sea demersal fish fauna of the Northern Gulf of Mexico. Journal of NorthWest Atlantic Fisheries Science 31: 1933.Google Scholar
Preston, G.L., Mead, P.D., Chapman, L.B., & Taumaia, P. (1999). Deep-bottom Fishing Techniques for the Pacific Islands: A Manual for Fishermen. Secretariat of the Pacific Community. Noumea, New Caledonia: Secretariat of the Pacific Community.Google Scholar
Priede, I.G. (1983). Use of satellites in marine biology. In: MacDonald, A.G. & Priede, I.G. (1983). Experimental Biology at Sea, pp. 350. New York: Academic Press.Google Scholar
Priede, I.G. (1984). A basking shark (Cetorhinus maximus) tracked by satellite together with simultaneous remote sensing. Fisheries Research 2: 201216.Google Scholar
Priede, I.G. (1985). Metabolic scope in fishes. In: Tytler, P. & Calow, P. (eds). Fish Energetics: New Perspectives , pp. 3364. London: Croom Helm.CrossRefGoogle Scholar
Priede, I.G. (1992). Wildlife telemetry: an introduction. In: Priede, I.G. & Swift, S.M. (eds.) Wildlife Telemetry: Remote Monitoring and Tracking of Animals, pp. 325. Chichester: Ellis Horwood.Google Scholar
Priede, I.G. (2013). Biogeography of the Oceans: a Review of Development of Knowledge of Currents, Fronts and Regional Boundaries from Sailing Ships in the Sixteenth Century to Satellite Remote Sensing. Pure and Applied Geophysics 171(6): 10131027. doi: 10.1007/s00024-013–0708-4CrossRefGoogle Scholar
Priede, I.G. (2014). Biogeography of the oceans: a review of development of knowledge of currents, fronts and regional boundaries from sailing ships in the sixteenth century to satellite remote sensing. Pure and Applied Geophysics 171: 10131027. DOI 10.1007/s00024-013-0708-4Google Scholar
Priede, I.G. & Bagley, P.M. (2000). In situ studies on deep-sea demersal fishes using autonomous unmanned lander platforms. Oceanography & Marine Biology, Annual Review 38: 357392.Google Scholar
Priede, I.G., Bagley, P., Armstrong, J.D., Smith, K.L., & Merrett, N.R. (1991). Direct Measurement of Active dispersal of Food-falls by abyssal demersal fishes. Nature 351: 647649.Google Scholar
Priede, I.G., Bagley, P.M., Smith, A., Creasey, S., & Merrett, N.R. (1994). Scavenging deep demersal fishes of the Porcupine Seabight (NE Atlantic Ocean); observations by baited camera, trap and trawl. Journal of the Marine Biological Association of the UK 74: 481498.Google Scholar
Priede, I.G., Bergstad, O.A., Miller, P.I., Vecchione, M., Gebruk, A., Falkenhaug, T., Billett, D.S.M., Craig, J., Dale, A.C., Shields, M.A., Tilstone, G.H., Sutton, T.T., Gooday, A.J., Inall, M.E., Jones, D.O.B., Martinez-Vicente, V., Menezes, G.M., Niedzielski, T., Sigurðsson, Þ., Rothe, N., Rogacheva, A., Alt, C.H.S., Brand, T., Abell, R., Brierley, A.S., Cousins, N.J., Crockard, D., Hoelzel, A.R., Høines, Å., Letessier, T.B., Read, J.F., Shimmield, T., Cox, M.J., Galbraith, J.K., Gordon, J.D.M., Horton, T., Neat, F., & Lorance, P., (2013). Does presence of a Mid Ocean Ridge enhance biomass and biodiversity? PLoS ONE 8(5): e61550. doi:10.1371/journal.pone.0061550CrossRefGoogle ScholarPubMed
Priede, I.G., Billett, D.S.M., Brierley, A.S., Hoelzel, A.R., Inall, M., Miller, P.I., Cousins, N.J., Shields, M.A., & Fujii, T. (2013). The ecosystem of the Mid-Atlantic ridge at the sub-polar front and Charlie Gibbs Fracture Zone; ECO-MAR project strategy and description of the sampling programme 2007–2010. Deep-Sea Research II 98: 220230. http://dx.doi.org/10.1016/j.dsr2.2013.06.012Google Scholar
Priede, I.G. & Froese, R. (2013). Colonisation of the deep-sea by fishes. Journal of Fish Biology 83: 15281550. doi:10.1111/jfb.12265CrossRefGoogle ScholarPubMed
Priede, I.G., Froese, R., Bailey, D.M., Bergstad, O.A., Collins, M.A., Dyb, J.E., Henriques, C., Jones, E.G. & King, N. (2006). The absence of sharks from abyssal regions of the world’s oceans. Proceedings of the Royal Society B. 273: 14351441. doi:10.1098/rspb.2005.3461Google Scholar
Priede, I.G., Godbold, J.A., King, N.J., Collins, M.A., Bailey, D.M., & Gordon, J.D.M. (2010). Deep-sea demersal fish species richness in the Porcupine Seabight, NE Atlantic Ocean: global and regional patterns. Marine Ecology 31: 247260. 10.1111/j.1439-0485.2009.00330.xGoogle Scholar
Priede, I.G., Godbold, J.A., Niedzielski, T., Collins, M.A., Bailey, D.M., Gordon, J.D.M., & Zuur, A.F. (2011). A review of the spatial extent of fishery effects and species vulnerability of the deep-sea demersal fish assemblage of the Porcupine Seabight, Northeast Atlantic Ocean (ICES Subarea VII). ICES Journal of Marine Science 68: 281289. doi:10.1093/icesjms/fsq045CrossRefGoogle Scholar
Priede, I.G. & Holliday, F.G.T. (1980). The use of tilting tunnel respirometer to investigate some aspects of metabolism and swimming activity of the plaice (Pleuronectes platessa L.). Journal of Experimental Biology 85: 295309.CrossRefGoogle Scholar
Priede, I.G. & Merrett, N.R. (1996). Estimation of abundance of abyssal demersal fishes; a comparison of data from trawls and baited cameras. Journal of Fish Biology 49(Suppl. A): 207216. doi: 10.1111/j.1095–8649.1996.tb06077.xCrossRefGoogle Scholar
Priede, I.G. & Smith, K.L. Jr. (1986). Behaviour of the abyssal grenadier, Coryphaenoides yaquinae, monitored using ingestible acoustic transmitters in the Pacific Ocean. Journal of Fish Biology 29(Suppl. A): 199206. doi: 10.1111/j.1095–8649.1986.tb05011.xCrossRefGoogle Scholar
Priede, I.G., Smith, K.L. Jr. & Armstrong, J.D. (1990). Foraging behaviour of abyssal grenadier fish: inferences from acoustic tagging and tracking in the North Pacific Ocean. Deep-sea Research 37: 81101.CrossRefGoogle Scholar
Priede, I.G., Williams, L.M., Wagner, H.-J., Thom, A, Brierley, I., Collins, M.A., Collin, S.P., Merrett, N.R., & Yau, C. (1999). Implication of the visual system in regulation of activity cycles in the absence of solar light: 2-[125I] iodomelatonin binding sites and melatonin receptor gene expression in the brains of demersal deep-sea gadiform fishes. Proceedings of the Royal Society. Series B 266: 22952302.Google Scholar
Prokofiev, A.M. (2007). Osteology and some other morphological characters of Howella sherborni, with a discussion of the systematic position of the genus (Perciformes, Percoidei). Journal of Ichthyology 47(6): 413426.Google Scholar
Prokofiev, A.M. (2014). Taxonomy and Distribution of Deepsea Herring (Bathyclupeidae) in Oceans. Journal of Ichthyology 54(8): 493500.CrossRefGoogle Scholar
Puig, P., Canals, M., Company, J.B., Martín, J., Amblàs, D., Lastras, G., Palanques, A., & Calafat, A.M. (2012). Ploughing the deep sea floor. Nature 489(7415): 286289.CrossRefGoogle ScholarPubMed
Pusceddu, A., Bianchellia, S., Martín, J., Puig, P., Palanques, A., Masqué, P., & Danovaro, R. (2014). Chronic and intensive bottom trawling impairs deep-sea biodiversity and ecosystem functioning. Proceedings of the National Academy of Sciences 111: 88618866.Google Scholar
Pusineri, C., Magnin, V., Meynier, L., Spitz, J., Hassani, S., & Ridoux, V. (2007). Food and feeding ecology of the common dolphin (Delphinus delphis) in the oceanic northeast atlantic and comparison with its diet in neritic areas. Marine Mammal Science 23: 3047.CrossRefGoogle Scholar
Pusineri, C., Vasseur, Y., Hassani, S., Meynier, L., Spitz, J., & Ridoux, V. (2005). Food and feeding ecology of juvenile albacore, Thunnus alalunga, off the Bay of Biscay: a case study. ICES Journal of Marine Science 62: 116122. doi:10.1016/j.icesjms.2004.09.004Google Scholar
Quattrini, A.M. & Demopoulos, A.W.J. (2016). Ectoparasitism on deep-sea fishes in the western North Atlantic: In situ observations from ROV surveys. International Journal for Parasitology: Parasites and Wildlife 5: 217228.Google Scholar
Quéro, J.-C., Du Buit, M.-H., & Vayne, J.-J. (1998). Les observations de poissons tropicaux et le réchauffement des eaux dans l’Atlantique Européen. Oceanologica Acta 21(2): 345351.CrossRefGoogle Scholar
Quéro, J.-C. & Ozouf-Costaz, C. (1991). Ostracoberyx paxtoni, nouvelle espèce des cötes est de l’Australie. Remarques sur les modifications morphologiques des Ostracoberyx au cors de leur croissance (Perciformes, Ostracoberycidae). Cybium 15(1): 4354.Google Scholar
Rabindranath, A., Daase, M., Falk-Petersen, S., Wold, A., Wallace, M.I., Berge, J. & Brierley, A.S. (2011). Seasonal and diel vertical migration of zooplankton in the High Arctic during the autumn midnight sun of 2008. Marine Biodiversity 41: 365382.CrossRefGoogle Scholar
Radchenko, O.A. (2015). The system of the suborder Zoarcoidei (Pisces, Perciformes) as inferred from molecular genetic data. Russian Journal of Genetics 51(11): 10961112.Google Scholar
Radchenko, O.A., Chereshnev, I.A., & Petrovskaya, A.V. (2010). Relationships and position of the genus Neozoarces of the subfamily Neozoarcinae in the system of the suborder Zoarcoidei (Pisces, Perciformes) by molecular–genetic data. Journal of Ichthyology 50(3): 246251.CrossRefGoogle Scholar
Radchenko, V.I. (2007). Mesopelagic fish community supplies ‘biological pump’. The Raffles Bulletin of Zoology: Supplement 14: 265271.Google Scholar
Rahmstorf, S. (2006). Thermohaline ocean circulation. In: Elias, S. A. (ed.) Encyclopedia of Quaternary Sciences. Amsterdam: Elsevier.Google Scholar
Raman, M. & James, P.S.B.R. (1990). Distribution and abundance of Lanternfishes of the family Myctophidae in the EEZ of India. In: Proceedings of the first workshop on scientific results of FORV Sagar Sampada, 5–7 June 1989, Kochi. http://eprints.cmfri.org.in/5189/Google Scholar
Ramsing, N. & Gundersen, J. (2013). Seawater and gases: tabulated physical parameters of interest to people working with microsensors in marine systems. www.unisense.comGoogle Scholar
Rass, T. S. (1955). Deep-sea fishes of the Kurile-Kamchatka trench. Trudy Instituta Okeanologii, Akademiya Nauk SSSR (Proceedings of the Institute of Oceanology, Academy of Sciences, USSR) (12: 328–339. [In Russian]Google Scholar
Rass, T.S., Grigorasch, V.A. & Spanovskaya, V.D. (1982). Deep-sea bottom fishes caught on the 14th cruise of the R/V Akademik Kurchatov. Proceedings of the Institute of Oceanology, Academy of Sciences, USSR 100: 337347 (1975). Translation Series No. 29 Virginia Institute of Marine Science, Gloucester Point, Virginia 23062, USAGoogle Scholar
Rees, J.-F., De Wergifosse, B., Noiset, O., Dubuisson, M., Janssens, B., & Thompson, E.M. (1998). The origins of marine bioluminescence: turning oxygen defence mechanisms into deep-sea communication tools. Journal of Experimental Biology 201: 12111221.CrossRefGoogle ScholarPubMed
Regan, C.T. (1925). Dwarfed males parasitic on the females in oceanic angler-fishes (Pediculati, Ceratioidea). Proceedings of the Royal Society of London B 97: 386400.Google Scholar
Regan, C.T. & Trewavas, E. (1932). Deep-sea angler fishes (Ceratioidea). The Carlsberg Foundation’s Oceanographical Expedition Round the World 1928–30 and Previous ‘Dana’ Expeditions under the Leadership of the Late Professor Johannes Schmidt. Dana Report. 2:1–113, Pls. 1–10.Google Scholar
Reichart, G. J., Lourens, L. J., & Zachariasse, W. J. (1998). Temporal variability in the northern Arabian Sea Oxygen Minimum Zone (OMZ) during the last 225,000 years. Paleoceanography 13: 607621.CrossRefGoogle Scholar
Reif, W.-E. (1985). Functions of scales and photophores in mesopelagic luminescent sharks. Acta Zoologica (Stockh.) 66: 111118.Google Scholar
Rex, M. & Etter, R. (2010). Deep-Sea Biodiversity: Pattern and Scale. Cambridge MA: Harvard University Press.Google Scholar
Rex, M. A., Etter, R. J., Morris, J. S., Crouse, J., McClain, C. R., Johnson, N. A., Stuart, C. T., Deming, J. W., Thies, R. & Avery, R. (2006). Global bathymetric patterns of standing stock and body size in the deep-sea benthos. Marine Ecology Progress Series 317: 18.CrossRefGoogle Scholar
Ribas, D., Muñoz, M., Casadevall, M., &, Gil de Sola, L. (2006). How does the northern Mediterranean population of Helicolenus dactylopterus dactylopterus resist fishing pressure? Fisheries Research 79(2006): 285293.Google Scholar
Rice, A.L. (1986). British Oceanographic Vessels 1800–1950. London: The Ray Society.Google Scholar
Rice, A.L., Aldred, R.G., Darlington, E., & Wild, R.A. (1982). The quantitative estimation of the deep-sea megabenthos: a new approach to an old problem. Oceanologica Acta 5: 6372.Google Scholar
Rigby, S. & Milsom, C.V. (2000). Origins, evolution, and diversification of zooplankton. Annual Review of Ecology, Evolution, and Systematics 31: 293313.Google Scholar
Rink, H. (1852). De danske Handelsdistricketer I Nordgrønland, deres geografiske Beskanffendhed og productive. Forste Deel. In : Grønland geografisk og statistisk beskrevet. 1 (1857)Google Scholar
Risso, A. (1810). Ichthyologie de Nice, ou histoire naturelle des poisons du Departement des Alpes Maritimes. Paris: Schoell.Google Scholar
Risso, A. (1820). Mémoire sur quelques poissons observes dans la mer de Nice. Journal De Physique, De Chimie, D'histoire Naturelle et Des Arts 91: 241255.Google Scholar
Risso, A. (1826). Histoire naturelle des principales productions de l’Europe Méridionale et particulièrement de celles des environs de Nice et des Alpes Maritimes; volume 3 Paris: Levrault.Google Scholar
Risso, A. & Poiteau, A. (1818–1822). Histoire naturelle des Orangiers. Paris: Audot.Google Scholar
Rittmeyer, E.N., Allen, A., Gründler, M.C, Thompson, D.K., & Austin, C.C. (2012). Ecological guild evolution and the discovery of the world’s smallest vertebrate. PLoS ONE (Public Library of Science) 7(1): e29797. doi:10.1371/journal.pone.0029797. PMC 3256195. PMID 22253785. Retrieved 11 January 2012.CrossRefGoogle ScholarPubMed
Roa-Varón, A. & Ortí, G. (2009). Phylogenetic relationships among families of Gadiformes (Teleostei, Paracanthopterygii) based on nuclear and mitochondrial data. Molecular Phylogenetics and Evolution 52: 688704.Google Scholar
Robalo, J.I., Sousa-Santos, C., Cabral, H., Castilho, R., & Almada, V.C. (2009). Genetic evidence fails to discriminate between Macroramphosus gracilis Lowe 1839 and Macroramphosus scolopax Linnaeus 1758 in Portuguese waters. Marine Biology 156: 17331737. doi: 10.1007/s00227-009–1197-yGoogle Scholar
Robbins, E.I., Porter, K.G., & Haberyan, K.A. (1985). Pellet microfossils: possible evidence for metazoan life in early Proterozoic time. Proceedings of the National Academy of Sciences USA 82: 58095813.CrossRefGoogle ScholarPubMed
Roberts, C.D. & Gomon, M.F. (2012). A review of giant roughies of the genus Hoplostethus (Beryciformes, Trachichthyidae), with descriptions of two new Australasian species. Memoirs of Museum Victoria 69: 341354.Google Scholar
Robertson, D.A. (1977). Planktonic eggs of the lanternfish, Lampanyctodes hectoris (family Myctophidae). Deep-Sea Research 24: 849852.CrossRefGoogle Scholar
Robertson, D.A. (1981). Possible functions of surface structure and size in some planktonic eggs of marine fishes. New Zealand Journal of Marine and Freshwater Research 15: 147153.CrossRefGoogle Scholar
Robins, C.H. & Martin, D.M. (1976). Haptenchelys texi. In: Robins, C.H. & Robins, C.R. (eds.) New Genera and Species of Dysommine and Synaphobranchine Eels (Synaphobranchidae) with an Analysis of the Dysomminae, pp. 267–274. Proceedings of the Academy of Natural Sciences of Philadelphia, 127: 249–280.Google Scholar
Robinson, A.R., Leslie, W.G., Theocharis, A., & Lascaratos, A. (2001). Mediterranean Sea Circulation. Encyclopedia of Ocean Sciences. San Diego, CA: Academic Press, 16891706.Google Scholar
Robinson, M. & Amemiya, C.T. (2014). Coelacanths. Current Biology 24(2): R62R63.CrossRefGoogle ScholarPubMed
Robison, B.H. (1972). Distribution of mid water fishes in the Gulf of Califonia. Copeia 1972(3): 448461.CrossRefGoogle Scholar
Robison, B.H. (1984). Herbivory by the myctophid fish Ceratoscopelus warmingii. Marine Biology 84: 119123.Google Scholar
Robison, B.H. (2000). The coevolutionof undersea vehicles and deep-sea research. Marine Technology Society Journal 33: 6573.Google Scholar
Robison, B.H. (2004). Deep pelagic biology. Journal of Experimental Marine Biology and Ecology 300: 253272.Google Scholar
Robison, B.H. & Reisenbichler, K.R. (2008). Macropinna microstoma and the paradox of its tubular eyes. Copeia 2008(4): 780784.CrossRefGoogle Scholar
Rodríguez-Cabello, C., González-Pola, C., & Sánchez, F. (2016). Migration and diving behaviour of Centrophorus squamosus in the NE Atlantic. Combining Electronic Tagging and Argo Hydrography to Infer Deep Ocean Trajectories. Deep-Sea Research I. 115: 4862.Google Scholar
Rodríguez-Cabello, C. & Sánchez, F. (2014). Is Centrophorus squamosus a highly migratory deep-water shark? Deep-Sea Research I 92: 110.CrossRefGoogle Scholar
Roe, H.S.J. (1988). Midwater biomass profiles over the Madeira Abyssal Plain and the contribution of copepods. Hydrobiologia 167/168: 169181.Google Scholar
Roe, H. S. J. & Badcock, J. (1984). The diel migrations and distributions within a mesopelagic community in the North East Atlantic. 5. Vertical migrations and feeding of fish. Progress in Oceanography 13: 389424.CrossRefGoogle Scholar
Roe, H. S. J. & Shale, D. M. (1979). A new multiple rectangular midwater trawl (RMT l+8M) and some modifications to the Institute of Oceanographic Sciences’ RMT 1+8. Marine Biology 50: 283288.CrossRefGoogle Scholar
Rogers, A.D. (2000). The role of the oceanic oxygen minima in generating biodiversity in the deep sea. Deep-Sea Research. II 47: 119148. doi:10.1016/S0967-0645(99)00107–1CrossRefGoogle Scholar
Rogers, A.D. (1994). The biology of seamounts. Advances in Marine Biology 30: 305350.Google Scholar
Rogers, A.D., Yesson, C., & Gravestock, P. (2015). A biophysical and economic profile of South Georgia and the South Sandwich Islands as Potential Large Scale Antarctic Protected areas. Advances in Marine Biology 70: 1286. doi: 10.1016/bs.amb.2015.06.001.CrossRefGoogle ScholarPubMed
Romer, A.S. (1970). The Vertebrate Body, 4th edition. Philadelphia: W.B. Saunders.Google Scholar
Rondelet, . (1554). Libri de piscibus marinus, in quibus verae piscium effigies expressae sunt. Lyons: Matthias Bonhomme.Google Scholar
Rosenblatt, R.H. & Butler, J.L. (1977). The ribbonfish genus Desmodema, with the description of a new species (Pisces, Trachipteridae) Fishery Bulletin 5: 843855.Google Scholar
Rosenblatt, R.H. & Cohen, D.M. (1986). Fishes living in deep sea thermal vents in the tropical eastern Pacific, with descriptions of a new genus and two new species of eelpouts (Zoarcidae). Transactions of the San Diego Society of Natural History 21: 7179.Google Scholar
Ross, L.G. & Gordon, J.D.M. (1978). Guanine and permeability in swimbladders of slope-dwelling fish. In: McLusky, D.S. et al. (eds.) Physiology and Behaviour of Marine Organisms: Proceedings of the 12th European Symposium on Marine Biology Stirling, Scotland, September 1977, pp. 113121. Oxford: Pergamon Press.CrossRefGoogle Scholar
Ross, S.W. & Quattrini, A.M. (2007). The fish fauna associated with deep coral banks off the southeastern United States. Deep-Sea Research I 54: 9751007.CrossRefGoogle Scholar
Ross, S.W., Quattrini, A.M., Roa-Varón, A.Y., & McClain, J.P. (2010). Species composition and distributions of mesopelagic fishes over the slope of the north-central Gulf of Mexico. Deep-Sea Research II 57: 19261956.CrossRefGoogle Scholar
Roule, L. (1913). Notice préliminaire sur Grimaldichthys profundissimus nov. gen. nov.sp. Possion abyssal recueilli à 6036 mètre de profondeur dans l’Ocean Atlantique par S.A.S le Prince de Monaco. Bulletin de l’Institut Oceanographique (Monaco) 261: 18.Google Scholar
Roule, M. L. (1922). Description de Scombrolabrax heterolepis nov. gen. nov. sp., poisson abyssal noveau de I’lle Madère. Bulletin de L’ Institut Océanographique, Monaco 408: 18.Google Scholar
Rountree, R.A., Juanes, F., Goudey, C.A., & Ekstrom, K.E. (2012). Is biological sound production important in the deep sea? In: Popper, A.N. & Hawkins, A. (eds.) The Effects of Noise on Aquatic Life, pp. 181183. New York: Springer Science & Business Media.Google Scholar
Rowat, D. & Gore, M. (2007). Regional scale horizontal and local scale vertical movements of whale sharks in the Indian Ocean off Seychelles. Fisheries Research 84:3240CrossRefGoogle Scholar
Rowden, A.A., Dower, J.F., Schlacher, T.A., Consalvey, M., & Clark, M.R. (2010). Paradigms in seamount ecology: fact, fiction and future. Marine Ecology 31: 226241.CrossRefGoogle Scholar
Rowe, G.T. (1983). Biomass and production of deep-sea macrobenthos. In: Rowe, G.T. (ed.) Deep-Sea Biology, pp. 97121. New York: Wiley-interscience.Google Scholar
Rowe, G.T. & Menzies, R. J. (1967). Use of sonic techniques and tension recordings as improvements in abyssal trawling. Deep-sea Research 14: 271274.Google Scholar
Rowe, G.T., Merrett, N., Shepherd, J., Needler, G., Hargrave, B., & Marietta, M. (1986). Estimates of direct transport of radioactive waste in the deep sea with species reference to organic carbon budgets. Oceanologica Acta 9: 199208.Google Scholar
Rowling, K. Hegarty, A., & Ives, M. (eds.). (2010). Ocean Perch (Helicolenus spp.). In: Status of Fisheries Resources in New South Wales, 2008/09, pp. 217220. Cronulla, NSW: Industry and Investment NSW.Google Scholar
Ruxton, G. D. & Houston, D. C. (2004). Energetic feasibility of an obligate marine scavenger. Marine Ecology Progress Series 266: 5963.Google Scholar
Ryther, J.H. (1956). Photosynthesis in the ocean as a function of light intensity. Limnology and Oceanography 1(1): 6170.CrossRefGoogle Scholar
Sabatés, A., Bozzano, A., & Vallvey, I. (2003). Feeding pattern and the visual light environment in myctophid fish larvae. Journal of Fish Biology 63: 14761490.Google Scholar
Sallan, L.C. & Coates, M.I. (2010). End-Devonian extinction and a bottleneck in the early evolution of modern jawed vertebrates. PNAS 107: 1013110135.Google Scholar
Sameoto, D., Wiebe, P., Runge, J., Postel, L., Dunn, J., Miller, C., & Coombs, S. (2000). Collecting zooplankton In: Harris, R. et al. (eds.) ICES Zooplankton Methodology Manual, pp. 5581. San Diego, CA: Academic Press.CrossRefGoogle Scholar
Sancho, G., Fisher, C.R., Mills, S., Micheli, F., Johnson, G.A., Lenihan, H.S., Peterson, C.H., & Mullineaux, L.S. (2005). Selective predation by the zoarcid fish Thermarces cerberus at hydrothermal vents. Deep-Sea Research I 52: 837844.Google Scholar
Sassa, C., Kawaguchi, K., Kinoshita, T., & Watanabe, C. (2002). Assemblages of vertical migratory mesopelagic fish in the transitional region of the western North Pacific. Fisheries Oceanography 11(4): 193204.CrossRefGoogle Scholar
Sato, K., Stewart, A.L., & Nakaya, K. (2013). Apristurus garricki sp. nov., a new deep-water catshark from the northern New Zealand waters (Carcharhiniformes: Scyliorhinidae). Marine Biology Research 9: 758767. http://dx.doi.org/10.1080/17451000.2013.765586Google Scholar
Sazonov, Y.I. (1996). Morphology and significance of the luminous organs in alepocephaloid fishes. In: Uiblein, F., Ott, J., & Stachowitsch, M. (eds.) Deep-Sea and Extreme Shallow-Water Habitats: Affinities and Adaptations. Biosystematics and Ecology Series 11:151163. Wien: Österreichische Akademie der Wissenschaften.Google Scholar
Schauer, J., Hissmann, K., & Fricke, H. (1997). A method for the deployment of externally attached, sonic fish tags from a manned submersible and their effects on coelacanths. Marine Biology 128: 359362.CrossRefGoogle Scholar
Schlee, S. (1973). A History of Oceanography; The Edge of an Unfamiliar World. London: Robert Hale.Google Scholar
Schmidt-Nielsen, K. (1987). Per Fredrik Thorkelsson Scholander, November 29, 1905–June 13, 1980. Biographical Memoir. Washington, DC: National Academy of Sciences.Google Scholar
Schmitter-Soto, J.J. (2008). The Oarfish, Regalecus glesne (Teleostei: Regalecidae), in the Western Caribbean. Caribbean Journal of Science 44: 125128.Google Scholar
Schmittner, A., Sarnthein, M.,.Kinkel, H., Bartoli, G., Bickert, T. et al. (2004). Global impact of the Panamanian Seaway closure. Eos 85(49): 526528.CrossRefGoogle Scholar
Schopf, T.J.M. (1980). Paleoceanography. Cambridge, MA: Harvard University Press.CrossRefGoogle Scholar
Schott, F.A. & McCreary, J.P. (2001). The monsoon circulation of the Indian Ocean. Progress in Oceanography 51: 1123.CrossRefGoogle Scholar
Schulte, P., Alegret, L., Arenillas, I., Arz, J.A., Barton, P.J. et al. (2010). The Chicxulub asteroid impact and mass extinction at the Cretaceous-Paleogene boundary. Science 327: 12141218. doi: 10.1126/science.1177265CrossRefGoogle ScholarPubMed
Scotese, C.R. (2002). PALEOMAP website. www.scotese.com, accessed 20 June 2014.Google Scholar
Scotese, C.R., Bambach, R.K., Barton, C., Van Der Voo, R., & Ziegler, A.M. (1979). Paleozoic base maps. The Journal of Geology 87: 217277.CrossRefGoogle Scholar
Sealy, T.S. (1974). Soviet Fisheries: A review. Paper 1075. Marine Fisheries Review 36(8): 522.Google Scholar
Searle, R. (2013). Mid-Ocean Ridges. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Sébert, P. (1997). Pressure effects on shallow-water fishes. In: Randall, D.J. & Farrell, A.P. (eds.) Deep-Sea Fishes, pp. 279323. San Diego, CA: Academic Press.CrossRefGoogle Scholar
Secombes, C.J. & Ellis, A.E. (2012). The immunology of teleosts. In: Roberts, R.J. (ed.) Fish Pathology, 4th edition, pp. 144166. London: Blackwell.Google Scholar
Seibel, B.A. & Drazen, J.C. (2007). The rate of metabolism in marine animals: environmental constraints, ecological demands and energetic opportunities. Philosophical Transactions of the Royal Society B 362: 20612078. doi:10.1098/rstb.2007.2101CrossRefGoogle ScholarPubMed
Seki, M.P. & Somerton, D.A. (1994). Feeding ecology and daily ration of the pelagic armorhead, Pseudopentaceros wheeleri, at Southeast Hancock Seamount. Environmental Biology of Fishes 39: 7384.Google Scholar
Shandikov, G.A., Eakin, R.R., & Usachev, S. (2013). Pogonophryne tronio, a new species of Antarctic short-barbeled plunderfish (Perciformes: Notothenioidei: Artedidraconidae) from the deep Ross Sea with new data on Pogonophryne brevibarbata. Polar Biology 36: 273289. doi: 10.1007/s00300-012–1258-4Google Scholar
Shao, K.-T., Iwamoto, T., Ho, H.-C., Cheng, T.-Y., & Chen, C.-Y. (2008). Species composition and distribution pattern of grenadiers (Family Bathygadidae, Macrouridae and Macrourididae) fromTaiwan. American Fisheries Society Symposium 63: 1729.Google Scholar
Sherrard, K.M. (2000). Cuttlebone morphology limits habitat depth in eleven species of Sepia (Cephalopoda: Sepiidae) Biological Bulletin 198: 404414.Google Scholar
Shibanov, V.N. & Vinnichenko, V.I. (2008). Russian investigations and the Fishery of Roundnose Grenadier in the North Atlantic. American Fisheries Society Symposium 63: 399412.Google Scholar
Shimokawa, T., Amaoka, K., Kajiwara, Y., & Suyama, S. (1995). Occurrence of Thalasenchelys coheni (Anguilliformes; Chlopsidae) in the West Pacific Ocean. Japanese Journal of Ichthyologgy 42: 8992.Google Scholar
Shinohara, G. & Imamura, H. (2007). Revisiting recent phylogenetic studies of ‘Scorpaeniformes’. Ichthyological Research 54: 9299. doi: 10.1007/s10228-006–0379-6CrossRefGoogle Scholar
Shreeve, R.S., Collins, M.A., Tarling, G.A., Main, C.E., Ward, P., & Johnston, N.M. (2009). Feeding ecology of myctophid fishes in the northern Scotia Sea. Marine Ecology Progress Series 386: 221236.CrossRefGoogle Scholar
Sidell, B.D. & O’Brien, K.M. (2006). When bad things happen to good fish: the loss of hemoglobin and myoglobin expression in Antarctic icefishes. Journal of Experimental Biology 209: 17911802.CrossRefGoogle ScholarPubMed
Siegelman-Charbit, L. & Planque, B. (2016). Abundant mesopelagic fauna at oceanic high latitudes. Marine Ecology Progress Series 546: 277282. doi: 10.3354/meps11661Google Scholar
Sielfeld, W.K., Vargas, M.F., & Fuenzalida, R.F. (1995). Peces mesopelágicos frente a la costa norte de Chile (18°25’-21°47’S). Investigaciones Marinas, Valparaiso 23: 8397.Google Scholar
Sigsbee, C.D. (1880). Deep-Sea Sounding and Dredging: A Description and Discussion of Methods and Appliances Used on Board the Coast and Geodetic Survey Steamer ‘Blake’. Washington, DC: Government Printing Office.Google Scholar
Sigurðsson, T., Kristinsson, K., Rätz, H-J., Nedreaas, K.H., Melnikov, S.P., & Reinert, J. (2006). The fishery for pelagic redfish (Sebastes mentella) in the Irminger Sea and adjacent waters. ICES Journal of Marine Science 63(4): 725736. doi: 10.1016/j.icesjms.2005.12.010CrossRefGoogle Scholar
Sigurðsson, T., Thorsteinsson, V. & Gustafsson, L. (2006). In situ tagging of deep-sea redfish: application of an underwater, fish-tagging system. ICES Journal of Marine Science 63: 523531.Google Scholar
Silverberg, N., Edenborn, H.M., Ouellet, G. & Beland, P. (1987). Direct evidence of a mesopelagic fish, Melanostigma atlanticum (Zoarcidae) spawning within bottom sediments. Environmental Biology of Fishes 20: 195202.Google Scholar
Sims, D.W., Southall, E.J., Richardson, A.J., Reid, P.C., & Metcalfe, J.D. (2003). Seasonal movements and behaviour of basking sharks from archival tagging: no evidence of winter hibernation. Marine Ecology Progress Series 248: 187196.CrossRefGoogle Scholar
Sinclair, E.H. & Stabeno, P.J. (2002). Mesopelagic nekton and associated physics of the southeastern Bering Sea. Deep-Sea Research II 49: 61276145.CrossRefGoogle Scholar
Skov, P.V. & Steffensen, J.F. (2003). The blood volumes of the primary and secondary circulatory system in the Atlantic cod Gadus morhua L., using plasma bound Evans Blue and compartmental analysis. Journal of Experimental Biology 206: 591599.Google Scholar
Smith, C.R. & Baco, A.R. (2003). Ecology of whale falls at the deep-sea floor. Oceanography and Marine Biology: An Annual Review 2003(41): 311354.Google Scholar
Smith, C.R., Glover, A.G., Treude, T., Higgs, N.D., & Amon, D.J. (2015). Whale-fall ecosystems: recent insights into ecology, paleoecology, and evolution. Annual Review of Marine Science 7: 571596. doi: 10.1146/annurev-marine-010213–135144Google Scholar
Smith, D.G. (2002). Saccopharyngidae, Swallower eels. In: Carpenter, K.E. (ed.) FAO Species Identification Guide for Fishery the Living Marine Resources of the Western Central Atlantic, Volume 2 Bony Fishes Part 1 (Acipenseridae to Grammatidae), pp. 758759. Rome: FAO. www.fao.org/3/a-y4161e/y4161e11.pdf accessed 3 March 2015.Google Scholar
Smith, J.L.B. (1939). A living fish of Mesozoic type. Nature 143: 455456. doi:10.1038/143455a0Google Scholar
Smith, K.L. Jr. (1978). Metabolism of the abyssopelagic rattail Coryphaenoides armatus measured in situ. Nature 274: 362364.CrossRefGoogle Scholar
Smith, K.L. Jr. (1982). Zooplankton of a bathyal benthic boundary layer: in situ rates of oxygen consumption and ammonium excretion. Limnology and Oceanography 27(3): 461471.CrossRefGoogle Scholar
Smith, K.L. Jr. (1987). Food energy supply and demand: A discrepancy between particulate organic carbon flux and sediment community oxygen consumption in the deep ocean. Limnology and Oceanography 32: 201220.CrossRefGoogle Scholar
Smith, K.L. Jr. & Hessler, R. R. (1974). Respiration of benthopelagic fishes: In situ measurements at 1230 metres. Science 184: 7273.Google Scholar
Smith, K.L. Jr., Kaufmann, R.S., Baldwin, R.J., & Carlucci, A.F. (2001) Pelagic-benthic coupling in the abyssal eastern North Pacific: An 8-year time-series study of food supply and demand. Limnology & Oceanography 46:543556.Google Scholar
Smith, K.L. Jr., Kaufmann, R.S., & Wakefield, W.W. (1993). Mobile megafaunal activity monitored with a time-lapse camera in the abyssal North Pacific. Deep-Sea Research I 40: 23072324.CrossRefGoogle Scholar
Smith, K.L. Jr., Ruhl, H.A., Bett, B.J., Billett, D.S.M., Lampitt, R.S., & Kaufmann, R.S. (2009). Climate, carbon cycling, and deep-ocean ecosystems. Proceedings of the National Academy of Sciences USA 106: 1921119218.CrossRefGoogle ScholarPubMed
Smith, K.L. Jr., Sherman, A.D., Huffard, C.L., McGill, P.R., Henthorn, R., Von Thun, S., Ruhl, H.A., Kahru, M., & Ohman, M.D. (2014). Large salp bloom export from the upper ocean and benthic community response in the abyssal northeast Pacific: Day to week resolution. Limnology & Oceanography 59: 745757.CrossRefGoogle Scholar
Smith, W.H.F. & Sandwell, D.T. (1997). Global seafloor topography from satellite altimetry and ship depth soundings. Science 277: 19561962.Google Scholar
Smith, W.L. & Wheeler, W.C. (2006). Venom evolution widespread in fishes: a phylogenetic road map for the bioprospecting of piscine venoms. Journal of Heredity 97: 206217.Google Scholar
Somero, G.N. (1992). Adaptations to high hydrostatic pressure. Annual Reviews of Physiology 54: 557577.Google Scholar
Somero, G.N. & SiebenaIler, J.F. (1979). Inefficient lactate dehydrogenases of deep-sea fishes. Nature 282: 100102.Google Scholar
Somiya, H. (1977). Bacterial bioluminescence in chlorophthalmid deep-sea fish: a possible interrelationship between the light organ and the eyes. Experientia 33: 906909.Google Scholar
Somiya, H., Yamakawa, T., & Okiyama, M. (1996). Okiyama Bathysauropis gigas, a deep-sea aulopiform fish with a peculiar iris process and a pure-cone retina. Journal of Fish Biology 49(Supplement A): 175181.Google Scholar
Soofiani, N.M. & Hawkins, A.D. (1982). Energetics costs at different levels of feeding in juvenile cod, Gadus morhua L. Journal of Fish Biology 21: 577592.CrossRefGoogle Scholar
Sorensen, P.W. & Wisenden, B.D. (eds.). (2015). Fish Pheromones and Related Cues. Chichester: Wiley Blackwell.Google Scholar
Sorenson, L., Santini, F., & Alfaro, M.E. (2014). The effect of habitat on modern shark diversification. Journal of Evolutionary Biology 27: 15361548.CrossRefGoogle ScholarPubMed
Springer, S. & Burgess, G.H. (1985). Two new dwarf dogsharks (Etmopterus, Squalidae), found off the Caribbean Coast of Colombia. Copeia 1985(3): 584591.CrossRefGoogle Scholar
Staiger, J. C. (1972). Bassogigas profundissimus (Pisces; Brotulidae) from the Puerto Rico Trench. Bulletin of Marine Science 22(l): 2633.Google Scholar
Star, B. & Jentoft, S. (2012). Why does the immune system of Atlantic cod lack MHC II? Bioessays 34: 648651.CrossRefGoogle ScholarPubMed
Star, B., Nederbragt, A.J., Jentoft, S., Grimholt, U., Malmstrøm, M. et al. (2011). The genome sequence of Atlantic cod reveals a unique immune system. Nature 477: 207210. doi:10.1038/nature10342CrossRefGoogle ScholarPubMed
Starr, R.M. & Green, K. (2010). Fishes of Las Gemelas Seamounts and Isla Del Coco: Preliminary Findings of September 2009 Submersible Surveys. San Diego, CA: Extension Publications University of California. https://escholarship.org/uc/item/9w01v9h0 Accessed 25 April 2015.Google Scholar
Staudigel, H., Hart, S.R., Pile, A., Bailey, B.E., Baker, E.T., Brooke, S., Connelly, D.P., Haucke, L., German, C.R., Hudson, I., Jones, D., Koppers, A.A.P., Konter, J., Lee, R., Pietsch, T.W., Tebo, B.M., Templeton, A.S., Zierenberg, R., & Young, C.M. (2006). Vailulu’u Seamount, Samoa: life and death on an active submarine volcano. Proceedings of the National Academy of Sciences of the USA 103: 64486453.Google Scholar
Stefanescu, C., Morales-Nin, B., & Perri, F. (1994). Fish assemblages on the slope in the Catalan Sea (Western Mediterranean): influence of a submarine canyon. Journal of the Marine Biological Association of the UK 74: 499512.Google Scholar
Stehmann, M. (1990). Rajidae. In: Quéro, J. C. et al. (eds.) Check-list of the Fishes of the Eastern Tropical Atlantic, vol. 1, pp. 2950. Lisbon, Portugal: Junta Nacional de Investigacao Cientifica e Tecnológica.Google Scholar
Stehmann, M. & Pompert, J. (2009). Bathyraja meridionalis. The IUCN Red List of Threatened Species 2009: e.T161619A5466024. http://dx.doi.org/10.2305/IUCN.UK.2009-2.RLTS.T161619A5466024.en. Downloaded on 22 February 2016.CrossRefGoogle Scholar
Stehmann, M.F.W. & Merrett, N.R. (2001). First records of advanced embryos and egg capsules of Bathyraja skates from the deep north-eastern Atlantic. Journal of Fish Biology 59: 338349.Google Scholar
Stein, D.L. (1980). Aspects of reproduction of liparid fishes from the continental slope and abyssal plain off Oregon, with notes on growth. Copeia 1980: 687699.Google Scholar
Stein, D.L. (1985). Towing large nets by single warp at abyssal depths: methods and biological results. Deep-Sea Research 32: 183200.Google Scholar
Stein, D.L. (2005). Descriptions of four new species, redescription of Paraliparis membranaceus, and additional data on species of the fish family Liparidae (Pisces, Scorpaeniformes) from the west coast of South America and the Indian Ocean. Zootaxa 1019: 125.CrossRefGoogle Scholar
Stein, D.L. & Chernova, N.V. (2009). Anatole Petrovich Andriashev (1910–2009) Copeia 2009(3): 628634. 2009.doi:10.1643/OT-09–099Google Scholar
Stein, D.L., Drazen, J.C., Schlining, K.L., Barry, J.P., & Kuhnz, L. (2006). Snailfishes of the central California coast: video, photographic and morphological observations. Journal of Fish Biology 69: 970–98.Google Scholar
Stein, D.L., Felley, J.D., & Vecchione, M. (2005). ROV observations of benthic fishes in the Northwind and Canada Basins. Arctic Ocean Polar Biology 28: 232237. doi:10.1007/s00300-004–0696-zGoogle Scholar
Stein, D.L. & Pearcy, W.G. (1982). Aspects of reproduction, early life history, and biology of macrourid fishes off Oregon, U.S.A. Deep-Sea Research. 29: 13131329.CrossRefGoogle Scholar
Stenseth, N.C., Mysterud, A., Ottersen, G., Hurrell, J.W., Chan, K.-S., & Lima, M. (2002). Ecological effects of climate fluctuations. Science 297: 12921296. doi: 10.1126/science.1071281CrossRefGoogle ScholarPubMed
Stergiou, K.I., Machias, A., Somarakis, S., & Kapantagakis, A. (2003). Can we define target species in Mediterranean trawl fisheries? Fisheries Research 59: 431435.CrossRefGoogle Scholar
Stevenson, D.E. & Kenaley, C.P. (2011). Revision of the manefish genus Paracaristius (Teleostei: Percomorpha: Caristiidae), with descriptions of a new genus and three new species. Copeia 2011(3): 385399.CrossRefGoogle Scholar
Stevenson, D.E. & Kenaley, C.P. (2013). Revision of the manefish genera Caristius and Platyberyx (Teleostei: Percomorpha: Caristiidae), with description of five new species. Copeia 2013(3): 415434.CrossRefGoogle Scholar
Stiassny, M.L.J. (1996). Basal Ctenosquamate relationships and the interrelationships of the myctophiform (Scopelomorph) fishes. In: Stiassny, M.L.J., Parenti, L.R., & Johnson, G.D. (eds.) Interrelationships of Fishes, pp. 405426. San Diego, CA: Academic Press.CrossRefGoogle Scholar
Stockton, W.L. & DeLaca, T.E. (1982). Food falls in the deep sea: occurrence, quality and significance. Deep-Sea Research 29: 157169.CrossRefGoogle Scholar
Stommel, H. & Arons, A.B. (1960). On the abyssal circulation of the world ocean- II. An idealized model of the circulation pattern and amplitude in oceanic basins. Deep Sea Research 6: 217218.Google Scholar
Stoner, A.W., Ryer, C.H., Parker, S.J., Auster, P.J., & Wakefield, W.W. (2008). Evaluating the role of fish behavior in surveys conducted with underwater vehicles. Canadian Journal of Fisheries and Aquatic Sciences 65: 12301243.CrossRefGoogle Scholar
Stowasser, G., Pond, D.W., & Collins, M.A. (2009). Using fatty acid analysis to elucidate the feeding habits of Southern Ocean mesopelagic fish. Marine Biology 156: 22892302. doi: 10.1007/s00227-009–1256-4Google Scholar
Stramma, L., Schmidtko, A., Levin, L.A., & Johnson, G.C. (2010). Ocean oxygen minima expansions and their biological impacts. Deep Sea Research Part I: Oceanographic Research Papers 57(4): 587595.CrossRefGoogle Scholar
Strange, K. (2016). Building a knowledge base for management of a new fishery: boarfish (Capros aper) in the Northeast Atlantic. Fisheries Research 174: 94102.CrossRefGoogle Scholar
Sturges, W. (2005). Deep-Water Exchange between the Atlantic, Caribbean and Gulf of Mexico. In: Circulation in the Gulf of Mexico: Observations and Models, pp. 263278. Geophysical Monograph Series 161. Washington, DC: American Geophysical Union. 10.1029/161GM019CrossRefGoogle Scholar
Suarez, S.S. (1975). The reproductive biology of Ogilbia cayorum, a viviparous brotulid fish. Bulletin of Marine Science 25: 143173.Google Scholar
Sudarto, Lalu X.C., Kosen, J.D., Tjakrawidjaja, A.H., Kusumah, R.V., Sadhotomo, B., Kadarusman, P.L., Slembrouck, J., & Paradis, E. (2010). Mitochondrial genomic divergence in coelacanths (Latimeria): slow rate of evolution or recent speciation? Marine Biology 157: 22532262. doi: 10.1007/s00227-010–1492-7CrossRefGoogle Scholar
Sulak, K.J. (1975). Talismania mekistonema, a new Atlantic species of the family Alepocephalidae (Pisces: Salmoniformes). Bulletin of Marine Science 25: 8893.Google Scholar
Sulak, K.J., Wenner, C.A., Sedberry, G.R., & Van Guelpen, L. (1985). Life history and systematics of deep-sea lizardfishes, genus. Bathysaurus (Synodontidae). Canadian Journal of Zoology 63: 623642.CrossRefGoogle Scholar
Summers, A.P., Hartel, K.E., & Koob, T.J. (1999). Agassiz, garman, albatross, and the collection of deep-sea fishes. Marine Fisheries Review 61: 5868.Google Scholar
Suntsov, A.V., Widder, E.A., & Sutton, T.T. (2008). Bioluminescence in larval fishes. In: Finn, R.N. & Kapoor, B.G. (eds.) Fish Larval Physiology, pp. 5188. Bergen, Norway: University of Bergen Press.Google Scholar
Sutton, T.T. (2005). Trophic ecology of the deep-sea fish Malacosteus niger (Pisces: Stomiidae): an enigmatic feeding ecology to facilitate a unique visual system? Deep-Sea Research I 52: 20652076.CrossRefGoogle Scholar
Sutton, T.T. (2013). Vertical ecology of the pelagic ocean: classical patterns and new perspectives. Journal of Fish Biology 83: 15081527.Google Scholar
Sutton, T.T., Clark, M.R., Dunn, D.C., Halpin, P.N., Rogers, A.D., Guinotte, J., Bograd, S.J., Angel, M.V., Perez, J.A.A., Wishner, K., Haedrich, R.L., Lindsay, D.J., Drazen, J.C., Vereshchakam, A., Piatkowski, U., Morato, T., Błachowiak-Samołyk, K., Robison, B.H., Gjerder, K.M., Pierrot-Bults, A., Bernalt, P., Reygondeau, G., & Heino, M. (2017). A global biogeographic classification of the mesopelagic zone. Deep-Sea Research I. doi.org/10.1016/j.dsr.2017.05.006Google Scholar
Sutton, T.T. & Hartel, K.E. (2004). New species of Eustomias (Teleostei: Stomiidae) from the Western North Atlantic, with a review of the subgenus Neostomias. Copeia 2004(1): 116121.Google Scholar
Sutton, T.T. & Hopkins, T.L. (1996a). Species composition, abundance, and vertical distribution of the stomiid (Pisces: Stomiiformes) fish assemblage of the Gulf of Mexico. Bulletin of Marine Science 59: 530–42.Google Scholar
Sutton, T.T. & Hopkins, T.L. (1996b). Trophic ecology of the stomiid (Pisces: Stomiidae) fish assemblage of the eastern Gulf of Mexico: strategies, selectivity and impact of a top mesopelagic predator group. Marine Biology 127: 179192.CrossRefGoogle Scholar
Sutton, T.T., Porteiro, F.M., Heino, M., Byrkjedal, I., Langhelle, G., Anderson, C.I.H., Horne, J., Søiland, H., Falkenhaug, T., Godø, O.R., & Bergstad, O.A. (2008). Vertical structure, biomass and topographic association of deep-pelagic fishes in relation to a mid-ocean ridge system. Deep-Sea Research II 55: 161184.Google Scholar
Sutton, T.T. & Sigurðsson, T. (2008). Vertical and horizontal distribution of mesopelagic fishes along a transect across the northern Mid-Atlantic Ridge. International Council for Exploration of the Sea CM 2008/C 16: 112.Google Scholar
Sutton, T.T., Wiebe, P.H., Madin, L.P., & Bucklin, A. (2010). Diversity and community structure of pelagic fishes to 5000 m depth in the Sargasso Sea. Deep-Sea Research II 57: 22202233.CrossRefGoogle Scholar
Suziumov, E.M. (1970–1979). Vessels, Scientific Research. The Great Soviet Encyclopedia, 3rd edition (1970–1979). New York: Macmillan.Google Scholar
Svendsen, F.M. & Byrkjedal, I. (2013). Morphological and molecular variation in Synaphobranchus eels (Anguilliformes: Synaphobranchidae) of the Mid-Atlantic Ridge in relation to species diagnostics. Marine Biodiversity 43: 407420. doi: 10.1007/s12526-013–0168-1Google Scholar
Sweetman, A.K., Smith, C.R., Dale, T., & Jones, D.O.B. (2014). Rapid scavenging of jellyfish carcasses reveals the importance of gelatinous material to deep-sea food webs. Proceedings of the Royal Society B 281: 20142210. http://dx.doi.org/10.1098/rspb.2014.2210Google Scholar
Swezey, R. R. & Somero, G. N. (1982). Polymerization thermodynamics and structural stabilities of skeletal muscle actins from vertebrates adapted to different temperatures and pressures. Biochemistry 21: 44964503.CrossRefGoogle ScholarPubMed
Swinney, G.N. (1994). Comments on the Atlantic species of the genus Evermannella (Scopelomorpha, Aulopiformes, Evermannellidae) with a re-evaluation of the status of Evermanella melanoderma. Journal of Fish Biology 44(5): 809819.Google Scholar
Tacon, A.G.J & Metian, M. (2015). Feed matters: satisfying the feed demand of aquaculture. Reviews in Fisheries Science & Aquaculture 23(1): 110. doi:10.1080/23308249.2014.987209Google Scholar
Takashima, R., Nishi, H., Huber, B.T., & Leckie, M. (2006). Greenhouse world and the Mesozoic ocean. Oceanography 19: 8292.Google Scholar
Talley, L.D., Lobanov, V., Ponomarev, V., Salyuk, A., Tishchenko, P., & Zhabin, I. (2003). Deep convection and brine rejection in the Japan Sea. Geophysical Research Letters 30(4): 1159.Google Scholar
Tamburini, C., Canals, M., Durrieu de Madron, X., Houpert, L., Lefèvre, D., et al. (2013). Deep-sea bioluminescence blooms after dense water formation at the ocean surface. PLoS ONE 8(7): e67523. doi:10.1371/journal.pone.0067523Google Scholar
Tandstad, M., Shotton, R., Sanders, J., & Carocci, F. (2011). Deep-sea fisheries. In: Review of the State of World Marine Fishery Resources, pp. 265277. FAO Fisheries and Aquaculture Technical Paper No. 569. Rome: FAO.Google Scholar
Tascheri, R., Saavedra-Nievas, J.C., & Roa-Uretad, R. (2010). Statistical models to standardize catch rates in the multi-species trawl fishery for Patagonian grenadier (Macruronus magellanicus) off Southern Chile. Fisheries Research 105: 200214.Google Scholar
Taylor, J.R. & Ferrari, R. (2011). Ocean fronts trigger high latitude phytoplankton blooms. Geophysical Research Letters 38, L23601.Google Scholar
Taylor, L.R., Compagno, L.J.V., & Struhsaker, P.J. (1983). Megamouth – a new species, genus, and family of lamnoid shark (Megachasma pelagios, family Megachasmidae) from the Hawaiian Islands. Proceedings of the California Academy of Sciences 43: 87110.Google Scholar
Tchernavin, V.V. (1953). Summary of the Feeding Mechanisms of a Deep Sea Fish, Chauliodus sloani Schneider. London: British Museum (Natural History).Google Scholar
Thacker, C.E. & Roje, D.M. (2009). Phylogeny of cardinalfishes (Teleostei: Gobiiformes: Apogonidae) and the evolution of visceral bioluminescence. Molecular Phylogenetics and Evolution 52: 735745.Google Scholar
The International Hydrographic Organisation. (2003). The History of GEBCO 1903–2003. Principality of Monaco: IHO.Google Scholar
Themelis, D.E. & Halliday, R.G. (2012). Species composition and relative abundance of the mesopelagic fish fauna in the slope sea off Nova Scotia. Northeastern Naturalist 19(2): 177200.CrossRefGoogle Scholar
Thiel, H. (1975). The size structure of the deep-sea benthos. Internationale Revue der gesamten Hydrobiologie und Hydrographie 60: 575606.Google Scholar
Thompson, E.M. & Rees, J.-F. (1995). Origins of luciferins: ecology of bioluminescence in marine fishes. Biochemistry and Molecular Biology of Fishes 4: 435466. doi:10.1016/S1873-0140(06)80021-4CrossRefGoogle Scholar
Thomsen, L. & McCave, I.N. (2000). Aggregation processes in the benthic boundary layer at the Celtic Sea continental margin. Deep-Sea Research I 47: 13891404.CrossRefGoogle Scholar
Thomson, A. (2003). The management of redfish (Sebastes mentella) in the north Atlantic Ocean - a stock in movement. In: Papers Presented at the Norway-FAO Expert Consultation on the Management of Shared Fish Stocks, pp. 192199. Bergen, Norway, 7–10 October 2002. FAO Fisheries Report. No. 695, Suppl. Rome: FAO.Google Scholar
Thurstan, R.H., Brockington, S., & Roberts, C.M. (2010). The effects of 118 years of industrial fishing on UK bottom trawl fisheries. Nature Communications 1: 15 doi: 10.1038/ ncomms1013CrossRefGoogle ScholarPubMed
Tighe, K.A. & McCosker, J.E. (2003). Two new species of the genus Chlopsis (Teleostei: Anguilliformes: Chlopsidae) from the southwestern Pacific. Zootaxa 236: 18.CrossRefGoogle Scholar
Torres, J. J., Belman, B.W., & Childress, J.J. (1979). Oxygen consumption rates of midwater fishes as a function of depth of occurrence. Deep-Sea Research. 26: 185197.CrossRefGoogle Scholar
Tota, B. (1978). Functional cardiac morphology and biochemistry in Atlantic Bluefin Tuna. In: Sharp, G.D. & Dizon, A.E. (eds.) The Physiological Ecology of Tunas, pp. 89112. New York: Academic Press.CrossRefGoogle Scholar
Tozer, H. & Dagit, D.D. (2004). Husbandry of spotted ratfish, Hydrolagus colliei. In: Smith, M., Warmolts, D., Thoney, D., & Hueter, R. (eds.) The Elasmobranch Husbandry Manual: Captive Care of Sharks, Rays and Their Relatives, pp. 487491. Columbus, OH: Ohio Biological Survey.Google Scholar
Tracey, D.M., Bull, B., Clark, M.R., & MacKay, K. (2004). Fish species composition on seamounts and adjacent slope in New Zealand waters. New Zealand Journal of Marine and Freshwater Research 38(1): 163182.Google Scholar
Trayanovsky, F.M. & Lisovsky, S.F. (1995). Russian (USSR) fisheries research in deep waters (below 500 m) in the North Atlantic. In: Hopper, A.G. (ed.) Deep-water Fisheries of the North Atlantic Slope, pp. 357365. Dordrecht, Netherlands: Kluwer.CrossRefGoogle Scholar
Treberg, J.R., and Speers-Roesch, B. (2016). Does the physiology of chondrichthyan fishes constrain their distribution in the deep sea? Journal of Experiment Biology 219: 615625.CrossRefGoogle ScholarPubMed
Tsukamoto, Y. (2002). Leptocephalus larvae of Pterothrissus gissu collected from the Kuroshio–Oyashio transition region of the western North Pacific, with comments on its metamorphosis. Ichthyological Research 49: 267269.Google Scholar
Tucker, G. H. (1951). Relation of fishes and other organisms to the scattering of underwater sound. Journal of Marine Research 10: 215238.Google Scholar
Tunnicliffe, V., Koo, B.F., Tyle, J. & So, S. (2010). Flatfish at seamount hydrothermal vents show strong genetic divergence between volcanic arcs. Marine Ecology 31: 158167.CrossRefGoogle Scholar
Tunnicliffe, V., Tyler, J., & Dower, J.F. (2013). Population ecology of the tongue fish Symphurus thermophilus (Pisces; Pleuronectiformes; Cynoglossidae) at sulphur-rich hydrothermal vents on volcanoes of the northern Mariana Arc. Deep-Sea Research II 92: 172182.CrossRefGoogle Scholar
Tuponogov, V.N., Orlov, A.M., & Kodolov, L.S. (2008). The most abundant grenadiers of the Russian Far East EEZ: distribution and basic biological patterns. American Fisheries Society Symposium 63: 285316.Google Scholar
Turner, J.R., White, E.M., Collins, M.A., Partridge, J.C., & Douglas, R.H. (2009). Vision in lanternfish (myctophidae): adaptations for viewing bioluminescence in the deep-sea. Deep-Sea Research-I 56: 10031017.Google Scholar
Tyco Electronics Corporation. (2012). Rochester Brand Engineered Cable Solutions for Harsh Environments.Google Scholar
Tyler, J.C. (1968). A monograph of Triacanthoidea. Academy of Natural Sciences Philadelphia Monographs 16: 1364.Google Scholar
Tytler, P. & Blaxter, J. H. S. (1973). Adaptation by cod and saithe to pressure changes. Netherlands Journal of Sea Research 7: 3145.CrossRefGoogle Scholar
Uiblein, F., Lorance, P., & Latrouite, D. (2003). Behaviour and habitat utilisation of seven fish species on the Bay of Biscay continental slope, NE Atlantic. Marine Ecology Progress Series 257: 223232.CrossRefGoogle Scholar
Uiblein, F., Nielsen, J.G. & Møller, P.R. (2008). Systematics of the ophidiid genus Spectrunculus (Teleostei: Ophidiiformes) with resurrection of S. crassus. Copeia 2008(5): 542551.Google Scholar
UN. (1982). United Nations Convention on the Law of the Sea. www.un.org/Depts/los/convention_agreements/texts/unclos/unclos_e.pdf Accessed 20 June 2014Google Scholar
UNESCO. (2009). Global Open Oceans and Deep Seabed (GOODS) – Biogeographic Classification. Paris: UNESCO-IOC. (IOC Technical Series, 84) 96pp.Google Scholar
Uyeno, T. & Tsutsumi, T. (1991). Stomach contents of Latimeria chalumnae and further notes on its feeding habits. Environmental Biology of Fishes 32: 275279.CrossRefGoogle Scholar
Vaillant, L. (1888). Expéditions scientifiques du Travailleur et du Talisman pendant les annees 1880, 1881, 1882, 1883. Paris: Masson.Google Scholar
van Denderen, P.D., van Kooten, T., & Rijnsdorp, A.D. (2013). When does fishing lead to more fish? Community consequences of bottom trawl fisheries in demersal food webs. Proceedings of the Royal Society. B 280: 20131883. http://dx.doi.org/10.1098/rspb.2013.1883CrossRefGoogle ScholarPubMed
Van der Grient, J.M.A. & Rogers, A.D. (2015). Body size versus depth: regional and taxonomical variation in deep-sea meio- and macrofaunal organisms. Advances in Marine Biology 71: 71108. http://dx.doi.org/10.1016/bs.amb.2015.07.002CrossRefGoogle ScholarPubMed
Van Der Laan, R., Eschmeyer, W.M., & Fricke, R. (2014). Family-group names of Recent fishes. Zootaxa 3882(2): 001230.Google Scholar
Van Dover, C.L. (2000). The Ecology of Deep-Sea Hydrothermal Vents. Princeton, NJ: Princeton University Press.CrossRefGoogle Scholar
van Ginneken, V., Antonissen, E., Müller, U.K., Booms, R., Eding, E., Verreth, J., & van den Thillart, G. (2005). Eel migration to the Sargasso: remarkably high swimming efficiency and low energy costs. Journal of Experimental Biology 208: 13291335.CrossRefGoogle Scholar
van Haren, H. (2004). Spatial variability of deep-ocean motions above an abyssal plain. Journal of Geophysical Research 109: C12014. doi:10.1029/2004JC002558.CrossRefGoogle Scholar
Vardaro, M.F., Parmley, D., & Smith, K.L. Jr. (2007). A study of possible ‘reef effects’ caused by a long-term time-lapse camera in the deep North Pacific. Deep-Sea Research. I 54: 12311240.CrossRefGoogle Scholar
Velasco, E.M., González, F., Amez, M., Méndez, E., & Punzón, A. (2010). First occurrence of the deepwater scorpionfish Setarches guentheri (Scorpaeniformes: Setarchidae) in Cantabrian waters: a northernmost occurrence in the eastern Atlantic. Marine Biodiversity Records 3: e82. doi:10.1017/S1755267210000722.CrossRefGoogle Scholar
Venter, P., Timm, P., Gunn, G., le Roux, E., Serfontein, C., Smith, P., Smith, E., Bensch, M., Harding, D., & Heemstra, P. (2000) Discovery of a viable population of coelacanths (Latimeria chalumnae Smith, 1939) at Sodwana Bay, South Africa. South African Journal of Science 96(11/12): 567568.Google Scholar
Videler, J.J. (1993). Fish Swimming. London: Chapman and Hall.CrossRefGoogle Scholar
Vitale, S., Ragonese, S., Cannizzaro, L., Fiorentino, F., & Mazzola, S. (2014). Evidence of trawling impact on Hoplostethus mediterraneus in the central–eastern Mediterranean Sea. Journal of the Marine Biological Association of the UK 94(3): 631640.CrossRefGoogle Scholar
VLIZ. (2009). Longhurst Biogeographical Provinces. Available online at www.vliz.be/vmdcdata/vlimar/downloads.php. Consulted on 8 July 2014.Google Scholar
Wagner, H.-J. (2001a). Brain areas in abyssal demersal fishes. Brain Behavior and Evolution 57: 301316.CrossRefGoogle ScholarPubMed
Wagner, H.-J. (2001b). Sensory brain areas in mesopelagic fishes. Brain Behavior and Evolution 57: 117133.Google Scholar
Wagner, H.-J. (2002). Sensory brain areas in three families of deep-sea fish (slickheads, eels, and grenadiers): comparison of mesopelagic and demersal species. Marine Biology 141: 807817.CrossRefGoogle Scholar
Wagner, H.-J. (2003). Volumetric analysis of brain areas indicates a shift in sensory orientation during development in the deep-sea grenadier Coryphaenoides armatus. Marine Biology 142: 791797.CrossRefGoogle Scholar
Wagner, H.-J., Douglas, R.H., Frank, T.M., & Roberts, N.W. (2009). A novel vertebrate eye using both refractive and reflective optics. Current Biology 19: 108114.CrossRefGoogle ScholarPubMed
Wagner, H.-J., Fröhlich, E., Negishi, K., & Collin, S.P. (1998). The eyes of deep-sea fish II. Functional morphology of the retina. Progress in Retinal and Eye Research 17: 637685.Google Scholar
Wagner, H.-J., Kemp, K., Mattheus, U., & Priede, I.G. (2007). Rhythms at the bottom of the deep sea: cyclic current flow changes and melatonin patterns in two species of demersal fish. Deep-Sea Research I 54: 19441956.CrossRefGoogle Scholar
Wahlberg, M., Westerberg, H., Aarestrup, K., Feunteun, E., Gargan, P., & Righton, D. (2014). Evidence of marine mammal predation of the European eel (Anguilla anguilla L.) on its marine migration. Deep-Sea Research I 86: 3238.CrossRefGoogle Scholar
Wakai, N., Takemura, K., Morita, T., & Kitao, A. (2014). Mechanism of deep-sea fish a-actin pressure tolerance investigated by molecular dynamics simulations. PLoS ONE 9(1): e85852. doi:10.1371/journal.pone.0085852Google Scholar
Wall, C.C., Rountree, R.A., Pomerleau, C., & Juanes, F. (2014). An exploration for deep-sea fish sounds off Vancouver Island from the NEPTUNE Canada ocean observing system. Deep-Sea Research I 83: 5764.Google Scholar
Wang, J.T.-M. & Chen, C.-T. (2001). A review of lanternfishes (Families: Myctophidae and Neoscopelidae) and the distributions around Taiwan and the Tungsha Islands with notes on seventeen new records. Zoological Studies 40: 103126.Google Scholar
Wardle, C.S. (1975). Limit of fish swimming speed. Nature (Lond.) 225: 725727. doi:10.1038/255725a0CrossRefGoogle Scholar
Wardle, C.S. (1983). Fish reactions to towed fishing gears. In Priede, I.G. & MacDonald, A.G. (eds.) Experimental Biology at Sea, pp. 168195. London: Academic Press.Google Scholar
Wardle, C.S., Tetteh-Lartey, N., Macdonald, A.G., Harper, A.A., & Penneq, J.-P. (1987). The effect of pressure on the lateral swimming muscle of the European eel Anguilla anguilla and the deep sea eel Histiobranchus bathybius; results of Challenger Cruise 6B/85. Comparative Biochemistry & Physiology 88A: 595598.CrossRefGoogle ScholarPubMed
Watanabe, H., Fujikura, K., Kojima, S., Miyazaki, J. I., & Fujiwara, Y. (2010). Japan: vents and seeps in close proximity. In: Kiel, S. (ed.) The Vent and Seep Biota: Aspects from Microbes to Ecosystems, pp. 379401. Netherlands: Springer. (doi:10.1007/978-90-481–9572-5_12)CrossRefGoogle Scholar
Watson, W. (1996). Sternoptychidae: Hatchetfishes. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 268283. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Watson, W. & Sandknop, E.M. (1996). Notosudidae: Paperbones. In: Moser, H.G. (ed.) Early Stages of Fishes in the California Current Region. California Cooperative Oceanic Fisheries Investigations, pp. 344347. Atlas No 33. La Jolla, CA: South West Fisheries Centre.Google Scholar
Weaver, P.P.E., Thomson, J., & Hunter, P.M. (1987). Introduction. In: Weaver, P.P.E. & Thomson, J. (eds.) Geology and Geochemistry of Abyssal Plains, pp. viixii. Geological Society Special Publication No 31. Oxford: Blackwell.Google Scholar
Webb, P. W. (1971). The swimming energetics of trout. II. Oxygen consumption and swimming efficiency. Journal of Experimental Biology 55: 521540.Google Scholar
Webb, T.J., Vanden, Berghe E., & O’Dor, R. (2010). Biodiversity’s big wet secret: the global distribution of marine biological records reveals chronic under-exploration of the deep pelagic ocean. PLoS ONE 5(8): e10223. doi:10.1371/journal.pone.0010223CrossRefGoogle ScholarPubMed
Weber, R.E., Hourdez, S., Knowles, F., & Lallier, F. (2003). Hemoglobin function in deep-sea and hydrothermal-vent endemic fish: Symenchelis parasitica (Anguillidae) and Thermarces cerberus (Zoarcidae). Journal of Experimental Biology 206: 26932702.CrossRefGoogle ScholarPubMed
Wegner, N.C., Sepulveda, C.A., Bull, K.B., & Graham, J.B. (2010). Gill morphometrics in relation to gas transfer and ram ventilation in high-energy demand teleosts: scombrids and billfishes. Journal of Morphology 71: 3649.Google Scholar
Wegner, N.C., Snodgrass, O.E., Dewar, H., & Hyde, J.R. (2015). Whole-body endothermy in a mesopelagic fish, the opah, Lampris guttatus. Science 348(6236): 786789. doi: 10.1126/science.aaa8902.Google Scholar
Wei, C.-L., Rowe, G.T., Escobar-Briones, E., Boetius, A., Soltwedel, T., et al. (2010). Global patterns and predictions of seafloor biomass using random forests. PLoS ONE 5(12): e15323. doi:10.1371/journal.pone.0015323Google Scholar
Wells, H.G., Huxley, J., & Wells, G.P. (1931). The Science of Life. London: Cassel.Google Scholar
Wessel, J.H. & Johnson, R.K. (1995, 1998). The Sternoptychinae of the Somali Current region of the western Indian ocean: An introduction to Somali current mesopelagic fish studies. In: Pierrot-Bults, A.C. & van der Spoel, S. (eds.) Pelagic Biogeography, pp. 372379. IOC Workshop Report No. 142 Paris: UNESCO.Google Scholar
Wessel, P. (2001). Global distribution of seamounts inferred from gridded Geosat/ERS-1 altimetry. Journal of Geophysical Research B106: 1943119442.CrossRefGoogle Scholar
Wessel, P., Sandwell, D.T., & Kim, S.-S. (2010). The global seamount census. Oceanography 23: 2433.Google Scholar
Westbrook, G.K. & Reston, T.J (2002). The accretionary complex of the Mediterranean ridge: tectonics, fluid flow and the formation of brine lakes. Marine Geology 186: 18.CrossRefGoogle Scholar
Wetherbee, B.M. (2000). Assemblage of deep-sea sharks on Chatham Rise, New Zealand. Fishery Bulletin 98: 189198.Google Scholar
White, W.T., Fahmi, Dharmadi, Potter, I.C. (2006). Preliminary investigation of artisanal deep-sea chondrichthyan fisheries in Eastern Indonesia. In: Shotton, R. (ed.). Deep Sea 2003: Conference on the Governance and Management of Deep-sea Fisheries. Part 2: Conference poster papers and workshop papers, pp. 381387. FAO Fisheries Proceedings 3/2, Rome: FAO.Google Scholar
Widder, E.A. (1998). A predatory use of counterillumination by the squaloid shark, Isistius brasiliensis. Environmental Biology of Fishes 53(3): 267273. doi:10.1023/A:1007498915860CrossRefGoogle Scholar
Widder, E.A., Robison, B.H., Reisenbichler, K.R., & Haddock, S.H.D. (2005). Using red light for in situ observations of deep-sea fishes. Deep-Sea Research I 52: 20772085.Google Scholar
Wiebe, P.H., Morton, A.W., Bradley, A.M. Backus, R.H., Craddock, J.E., Barber, V., Cowles, T.J., & Flier, G.R. (1985). New developments in the MOCNESS, an apparatus for sampling zooplankton and micronekton. Marine Biology 87: 313323.CrossRefGoogle Scholar
Williams, A., Koslow, J.A., & Last, P.R. (2001). Diversity, density and community structure of the demersal fish fauna of the continental slope off Western Australia (20 to 35° S). Marine Ecology Progress Series 212: 247263.CrossRefGoogle Scholar
Williams, A., Last, P.R., Gomon, M.F., & Paxton, J.R. (1996). Species composition and checklist of the demersal ichthyofauna of the continental slope off Western Australia (20–35°5) Records of the Western Australian Museum 18: 135155.Google Scholar
Willughby, F. (1686). De Historia Piscium Libri Quatuor. Oxford.Google Scholar
Wilson, G. D. F. (1999). Some of the deep-sea fauna is ancient. Crustaceana 72: 10201030.Google Scholar
Wilson, R.R. & Waples, R.S. (1984), Electrophoretic and biometric variability in the abyssal grenadier Coryphaenoides (N.) armatus of the western North Atlantic, eastern Pacific and Eastern North Pacific Oceans. Marine Biology 80: 227237.Google Scholar
Wippelhauser, G.S., Miller, M.J., & McCleave, J.D. (1996). Evidence of spawning and the larval distributions of snipe eels (family nemichthyidae) in the Sargasso Sea. Bulletin of Marine Science 59(2): 298309.Google Scholar
Wishner, K.F. (1980). The biomass of the deep-sea benthopelagic plankton. Deep-Sea Research Part A. Oceanographic Research Papers 27(3–4): 205216.Google Scholar
Wittmann, A.C. & Pörtner, H.-O. (2013). Sensitivities of extant animal taxa to ocean acidification. Nature Climate Change. 3: 9951001 doi:10.1038/nclimate1982Google Scholar
Wolff, T. (1961). The deepest recorded fishes. Nature 190: 283284.Google Scholar
Wolff, T. (1971). Archimede dive 7 to 4160 at Madeira: observations and collecting results. Videnskabelige Meddelelser dansk naturhistorisk Forening Kobenhavn 134: 127147.Google Scholar
Wolff, T. (2002). The Danish Dana Expedition, 1928–30: Purpose and accomplishments, mainly in the Indo-Pacific. In: Benson, K.R. & Rehbock, P.F. (eds.) Oceanographic History, the Pacific & Beyond, pp. 196203. Proceedings of the 5th International Congress on the History of Oceanography. Seattle: University of Washington Press.Google Scholar
Woodward, A.S. (1898). The antiquity of the deep-sea fish-fauna. Journal of Natural Science 12: 257260.Google Scholar
Wright, E.P. (1870). Notes on Sponges. 1, On Hyalonema mirabilis, Gray. 2, Aphrocallistes Bocagei sp. Nov. 3, On a new Genus and Species of Deep Sea Sponge. Quarterly Journal of Microscopical Science 10: 19, pls I-III.Google Scholar
Wright, J.J., Konwar, K.M., & Hallam, S.J. (2012). Microbial ecology of expanding oxygen minimum zones. Nature Reviews Microbiology 10: 381394.Google Scholar
Wynne-Edwards, V.C. (1962). Animal Dispersion in Relation to Social Behaviour. Edinburgh: Oliver & Boyd.Google Scholar
Wyville, T.C. (1873). The Depths of the Sea. London: Macmillan.Google Scholar
Yabe, M. (1991). Bolinia euryptera, a new genus and species of sculpin (Scorpaeniformes; Cottidae) from the Bering Sea. Copeia 1991(2): 329339. (Ref. 39612)Google Scholar
Yamaguchi, M. (2000). Phylogenetic analyses of myctophid fishes using morphological characters: progress, problems, and future perspectives. Japanese Journal of Ichthyology 2000(47): 87107.Google Scholar
Yamamoto, J., Hirose, M., Ohtani, T., Sugimoto, K., Hirase, K., Shimamoto, N., Shimura, T., Honda, N., Fujimora, Y., & Mukai, T. (2008). Transportation of organic matter to the seafloor by carrion falls of the giant jellyfish Nemopilema nomurai in the Sea of Japan. Marine Biology 153: 311317.Google Scholar
Yamanoue, Y. & Yoseda, K. (2001). A new species of the genus Malakichthys (Perciformes: Acropomatidae) from Japan. Ichthyological Research 48: 257261.Google Scholar
Yanagimoto, T. & Humphreys, R.L. Jr. (2005). Maturation and reproductive cycle of female armorhead Pseudopentaceros wheeleri from the southern Emperor–northern Hawaiian Ridge Seamounts. Fisheries Science 71: 10591068.Google Scholar
Yancey, P.H., Gerringer, M.E., Drazen, J.C., Rowden, A.A., & Jamieson, A. (2014). Marine fish may be biochemically constrained from inhabiting the deepest ocean depths. Proceedings of National Academy of Science USA 111(12): 44614465. doi: 10.1073/pnas.1322003111Google Scholar
Yancey, P.H., Lawrence-Berrey, R., & Douglas, M.D. (1989). Adaptations in mesopelagic fishes. I. Buoyant glycosaminoglycan layers in species without diel vertical migrations. Marine Biology 103: 453459.Google Scholar
Yancey, P.H. & Siebenaller, J.F. (2015). Co-evolution of proteins and solutions: protein adaptation versus cytoprotective micromolecules and their roles in marine organisms. Journal of Experimental Biology 218: 18801896. doi:10.1242/jeb.114355CrossRefGoogle ScholarPubMed
Yang, J. & Huang, Z. (1986). The fauna and geographical distribution of deep-pelagic fishes in the South China Sea. In: Uyeno, T., Arai, R., Taniuchi, T., & Matsuura, K., (eds.) Indo-Pacific Biology: Proceedings of the 2nd International Conference on Indo-Pacific Fishes, pp. 461164. Sado, Japan: Ichthyologic Society of Japan.Google Scholar
Yau, C., Collins, M.A., & Everson, I. (2000), Commensalism between a liparid fish (Careproctus sp.) and stone crabs (Lithodidae) photgraphed in situ using a baited camera. Journal of the Marine Biological Society of the UK 80: 379380.Google Scholar
Ye, Y. & Cochrane, K. (2011), Global overview of marine fishery resources. In: Review of the state of world marine fishery resources, pp. 318. FAO Fisheries and Aquaculture Technical Paper No. 569. Rome: FAO.Google Scholar
Yeh, J. & Drazen, J.C. (2009). Depth zonation and bathymetric trends of deep-sea megafaunal scavengers of the Hawaiian Islands. Deep-Sea Research I 56: 251266.CrossRefGoogle Scholar
Yesson, C., Clark, M.R., Taylor, M., & Rogers, A.D. (2011). The global distribution of seamounts based on 30-second bathymetry data. Deep-Sea Research I 58(4): 442453.doi.org/10.1016/j.dsr.2011.02.004.CrossRefGoogle Scholar
Yoshinaga, T., Miller, M.J., Yokouchi, K., Otake, T., Kimura, S., Aoyama, J., Watanabe, S., Shinoda, A., Oya, M., Miyazaki, S., Zenimoto, K., Sudo, R., Takahashi, T., Ahn, H., Manabe, R., Hagihara, S., Morioka, H., Itakura, H., Machida, M., Ban, K., Shiozaki, M., Ai, B., & Tsukamoto, K. (2011). Genetic identification and morphology of naturally spawned eggs of the Japanese eel Anguilla japonica collected in the western North Pacific. Fisheries Science 77: 983992. doi: 10.1007/s12562-011–0418-8CrossRefGoogle Scholar
Zahl, P.A. (1953). Fishing in the whirlpool of Charybdis. National Geographic Magazine 104(5): 579618.Google Scholar
Zahuranec, B.J. (2000). Zoogeography and systematics of the lanternfishes of the genus Nannobrachium (Myctophidae: Lampanyctini). Smithsonian Contributions to Zoology 607: 169.Google Scholar
Zenkevitch, L.A. & Birstein, J.A. (1960). On the problem of the antiquity of the deep-sea fauna. Deep-Sea Research 7: 1023.Google Scholar
Zezina, O.N. (1997). Biogeography of the bathyal zone. Advances in Marine Biology 32: 389420.CrossRefGoogle Scholar
Zintzen, V., Roberts, C.D., Anderson, M.J., Stewart, A.L., Struthers, C.D., & Harvey, E.S. (2011). Hagfish predatory behaviour and slime defence mechanism. Scientific Reports. 1, 131. doi:10.1038/srep00131 (2011).Google Scholar
Zintzen, V., Rogers, K.M., Roberts, C.D., Stewart, A.L., & Anderson, M.J. (2013). Hagfish feeding habits along a depth gradient inferred from stable isotopes. Marine Ecology Progress Series 485: 223234. doi: 10.3354/meps10341Google Scholar
Zylinski, S. & Johnsen, S. (2011). Mesopelagic cephalopods switch between transparency and pigmentation to optimize camouflage in the deep. Current Biology 21: 19371941. doi :10.1016/j.cub.2011.10.014Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure [email protected] is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Imants G. Priede, University of Aberdeen
  • Book: Deep-Sea Fishes
  • Online publication: 24 August 2017
  • Chapter DOI: https://doi.org/10.1017/9781316018330.010
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Imants G. Priede, University of Aberdeen
  • Book: Deep-Sea Fishes
  • Online publication: 24 August 2017
  • Chapter DOI: https://doi.org/10.1017/9781316018330.010
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Imants G. Priede, University of Aberdeen
  • Book: Deep-Sea Fishes
  • Online publication: 24 August 2017
  • Chapter DOI: https://doi.org/10.1017/9781316018330.010
Available formats
×