Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-27T06:54:24.027Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  09 March 2018

Roger T. Hanlon
Affiliation:
Marine Biological Laboratory, Massachusetts
John B. Messenger
Affiliation:
University of Cambridge
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
Cephalopod Behaviour , pp. 291 - 354
Publisher: Cambridge University Press
Print publication year: 2018

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

References

Systematists have an irritating habit of every now and then changing the names of animals. This is not our fault. However, it can lead to confusion: the giant Pacific octopus that many of us used to know as Octopus dofleini is now called Enteroctopus dofleini. A squid often referred to in this book as Loligo pleii has recently become Doryteuthis pleii.

In response to this problem we have decided, throughout this book, to use the generic name of an animal as it appears in the original published article.

Adamo, S. A. & Hanlon, R. T. (1996). Do cuttlefish (Cephalopoda) signal their intentions to conspecifics during agonistic encounters? Animal Behaviour, 52, 7381.CrossRefGoogle Scholar
Adamo, S. A. & Weichelt, K. J. (1999). Field observations of schooling in the oval squid, Sepioteuthis lessoniana (Lesson, 1830). Journal of Molluscan Studies, 65, 377380.CrossRefGoogle Scholar
Adamo, S. A., Brown, W. M., Kinig, A. J. et al. (2000). Agonistic and reproductive behaviours of the cuttlefish Sepia officinalis in a semi-natural environment. Journal of Molluscan Studies, 66, 417419.CrossRefGoogle Scholar
Adamo, S. A., Ehgoetz, K., Sangster, C. & Whitehorne, I. (2006). Signaling to the enemy? – Body pattern expression and its response to external cues during hunting in the cuttlefish Sepia officinalis (Cephalopoda). Biological Bulletin, 210, 192200.CrossRefGoogle Scholar
Agin, V., Chichery, R. & Chichery, M. P. (2001). Effects of learning on cytochrome oxidase activity in cuttlefish brain. Neuroreport, 12, 113116.CrossRefGoogle ScholarPubMed
Agin, V., Chichery, R., Chichery, M. P. et al. (2006). Behavioural plasticity and neural correlates in adult cuttlefish. Vie et Milieu – Life and Environment, 56, 8187.Google Scholar
Agin, V., Chichery, R., Maubert, E. & Chichery, M. P. (2003). Time-dependent effects of cycloheximide on long-term memory in the cuttlefish. Pharmacology Biochemistry and Behavior, 75, 141146.CrossRefGoogle ScholarPubMed
Agin, V., Dickel, L., Chichery, R. & Chichery, M. P. (1998). Evidence for a specific short-term memory in the cuttlefish, Sepia. Behavioural Processes, 43, 329334.CrossRefGoogle ScholarPubMed
Aitken, J. P., O’Dor, R. K. & Jackson, G. D. (2005). The secret life of the giant Australian cuttlefish Sepia apama (Cephalopoda): behaviour and energetics in nature revealed through radio acoustic positioning and telemetry (RAPT). Journal of Experimental Marine Biology and Ecology, 320, 7791.CrossRefGoogle Scholar
Akkaynak, D., Allen, J. J., Mathger, L. M., Chiao, C. C. & Hanlon, R. T. (2013). Quantification of cuttlefish (Sepia officinalis) camouflage: a study of color and luminance using in situ spectrometry. Journal of Comparative Physiology A: Neuroethology Sensory Neural and Behavioral Physiology, 199, 211225.CrossRefGoogle ScholarPubMed
Alcock, J. (2013). Animal Behavior: An Evolutionary Approach. Sunderland, MA: Sinauer Associates, Inc.Google Scholar
Aldred, R. G., Nixon, M. & Young, J. Z. (1978). The blind octopus, Cirrothauma. Nature, 275, 547549.CrossRefGoogle Scholar
Aldred, R. G., Nixon, M. & Young, J. Z. (1983). Cirrothauma murrayi Chun, a finned octopod. Philosophical Transactions of the Royal Society of London B, 301, 154.Google Scholar
Allcock, A. L., Cooke, I. R. & Strugnell, J. M. (2011). What can the mitochondrial genome reveal about higher-level phylogeny of the molluscan class Cephalopoda? Zoological Journal of the Linnean Society, 161, 573586.CrossRefGoogle Scholar
Allen, A., Michels, J. & Young, J. Z. (1985). Memory and visual discrimination by squids. Marine Behaviour and Physiology, 11, 271282.CrossRefGoogle Scholar
Allen, A., Michels, J. & Young, J. Z. (1986). Possible interactions between visual and tactile memories in Octopus. Marine Behaviour and Physiology, 12, 8197.CrossRefGoogle Scholar
Allen, J. J., Bell, G. R. R., Kuzirian, A. M. & Hanlon, R. T. (2013). Cuttlefish skin papilla morphology suggests a muscular hydrostatic function for rapid changeability. Journal of Morphology, 274, 645656.CrossRefGoogle ScholarPubMed
Allen, J. J., Bell, G. R. R., Kuzirian, A. M., Velankar, S. S. & Hanlon, R. T. (2014). Comparative morphology of changeable skin papillae in octopus and cuttlefish. Journal of Morphology, 275, 371390.CrossRefGoogle ScholarPubMed
Allen, J. J., Mäthger, L. M., Barbosa, A. & Hanlon, R. T. (2009). Cuttlefish use visual cues to control three-dimensional skin papillae for camouflage. Journal of Comparative Physiology A: Neuroethology Sensory Neural and Behavioral Physiology, 195, 547555.CrossRefGoogle ScholarPubMed
Allen, J. J., Mäthger, L. M., Barbosa, A. et al. (2010a). Cuttlefish dynamic camouflage: responses to substrate choice and integration of multiple visual cues. Proceedings of the Royal Society B – Biological Sciences, 277, 10311039.CrossRefGoogle ScholarPubMed
Allen, J. J., Mäthger, L. M., Buresch, K. C. et al. (2010b). Night vision by cuttlefish enables changeable camouflage. Journal of Experimental Biology, 213, 39533960.CrossRefGoogle ScholarPubMed
Alonso, M. K., Crespo, E. A., Garcia, N. A. et al. (2002). Fishery and ontogenetic driven changes in the diet of the spiny dogfish, Squalus acanthias, in Patagonian waters, Argentina. Environmental Biology of Fishes, 63, 193202.CrossRefGoogle Scholar
Altman, J. S. (1967). The behaviour of Octopus vulgaris Lam. in its natural habitat: a pilot study. Underwater Association Reports, 1966 –1976, 7783.Google Scholar
Alupay, J. S., Hadjisolomou, S. P. & Crook, R. J. (2014). Arm injury produces long-term behavioral and neural hypersensitivity in octopus. Neuroscience Letters, 558, 137142.CrossRefGoogle ScholarPubMed
Alves, C., Chichery, R., Boal, J. G. & Dickel, L. (2007). Orientation in the cuttlefish Sepia officinalis: response versus place learning. Animal Cognition, 10, 2936.CrossRefGoogle ScholarPubMed
Alves, C., Darmaillacq, A. S., Shashar, N. & Dickel, L. (2007). Field and laboratory observations of Sepia (Doratosepion) elongata, d’Orbigny, 1845. Veliger, 48, 313316.Google Scholar
Amaratunga, T. (1980). Preliminary estimates of predation by the short-finned squid (Illex illecebrosus) on the Scotian Shelf. North Atlantic Fisheries Organization Scientific Council Research Document, 80 /II/31, 113.Google Scholar
Amaratunga, T. (1983). The role of cephalopods in the marine ecosystem, in Advances in Assessment of World Cephalopod Resources (ed. Caddy, J. F.), pp. 379415. Rome: FAO.Google Scholar
Amaratunga, T. (1987). Population biology, in Cephalopod Life Cycles, Vol. II. Comparative Reviews (ed. Boyle, P. R.), pp. 239252. London: Academic Press.Google Scholar
Amaratunga, T., Roberge, M., Young, J. & Uozumi, Y. (1980). Summary of joint Canada/Japan research program on short-finned squid (Illex illecebrosus): emigration and biology. North Atlantic Fisheries Organization Scientific Council Research Document, 80 /11/40, No. N071, 120.Google Scholar
Ambrose, R. F. (1982). Shelter utilization by the molluscan cephalopod Octopus bimaculatus. Marine Ecology Progress Series, 7, 6773.CrossRefGoogle Scholar
Ambrose, R. F. (1983). Midden formation by octopuses: the role of biotic and abiotic factors. Marine Behaviour and Physiology, 10, 137144.CrossRefGoogle Scholar
Ambrose, R. F. (1984). Food preferences, prey availability, and the diet of Octopus bimaculatus Verrill. Journal of Experimental Marine Biology and Ecology, 77, 2944.CrossRefGoogle Scholar
Ambrose, R. F. (1986). Effects of octopus predation on motile invertebrates in a rocky subtidal community. Marine Ecology Progress Series, 30, 261273.CrossRefGoogle Scholar
Ambrose, R. F. (1988). Population dynamics of Octopus bimaculatus: influence of life history patterns, synchronous reproduction and recruitment. Malacologia, 29, 2339.Google Scholar
Ambrose, R. F. & Nelson, B. V. (1983). Predation by Octopus vulgaris in the Mediterranean. Marine Ecology, 4, 251261.CrossRefGoogle Scholar
Anderson, J. C., Baddeley, R. J., Osorio, D. et al. (2003). Modular organization of adaptive colouration in flounder and cuttlefish revealed by independent component analysis. Network-Computation in Neural Systems, 14, 321333.CrossRefGoogle ScholarPubMed
Anderson, J. R. (2005). Cognitive Psychology and Its Implications. New York: Worth.Google Scholar
Anderson, R. C. & Mather, J. A. (2007). The packaging problem: bivalve prey selection and prey entry techniques of the octopus Enteroctopus dofleini. Journal of Comparative Psychology, 121, 300305.CrossRefGoogle ScholarPubMed
Anderson, R. C., Hughes, P. D., Mather, J. A. & Steele, C. W. (1999). Determination of the diet of Octopus rubescens Berry, 1953 (Cephalopoda: Octopodidae), through examination of its beer bottle dens in Puget Sound. Malacologia, 41, 455460.Google Scholar
Anderson, R. C., Mather, J. A. & Sinn, D. L. (2008). Octopus senescence: forgetting how to eat clams. Festivus, 40, 5557.Google Scholar
Anderson, R. C., Sinn, D. L. & Mather, J. A. (2008). Drilling localization on bivalve prey by Octopus rubescens Berry, 1953 (Cephalopoda: Octopodidae). Veliger, 50, 326328.Google Scholar
Anderson, R. C., Wood, J. B. & Mather, J. A. (2008). Octopus vulgaris in the Caribbean is a specializing generalist. Marine Ecology Progress Series, 371, 199202.CrossRefGoogle Scholar
Anderson, T. J. (1997). Habitat selection and shelter use by Octopus tetricus. Marine Ecology Progress Series, 150, 137148.CrossRefGoogle Scholar
Anderson, T. J. (1999). Morphology and biology of Octopus maorum Hutton 1880 in northern New Zealand. Bulletin of Marine Science, 65, 676.Google Scholar
Andersson, M. (1982). Sexual selection, natural selection and quality advertisement. Biological Journal of the Linnean Society, 17, 375393.CrossRefGoogle Scholar
Andersson, M. (1994). Sexual Selection. Princeton: Princeton University Press.CrossRefGoogle Scholar
Andersson, M. & Simmons, L. W. (2006). Sexual selection and mate choice. Trends in Ecology and Evolution, 21, 296302.CrossRefGoogle ScholarPubMed
Andre, M., Johansson, T., Delory, E. & Van Der Schaar, M. (2007). Foraging on squid: the sperm whale mid-range sonar. Journal of the Marine Biological Association of the United Kingdom, 87, 5967.CrossRefGoogle Scholar
Andre, M., Sole, M., Lenoir, M. et al. (2011). Low-frequency sounds induce acoustic trauma in cephalopods. Frontiers in Ecology and the Environment, 9, 489493.CrossRefGoogle Scholar
Antonelis, G. A., Lowry, M. S., DeMaster, D. P. & Fiscus, C. H. (1987). Assessing northern elephant seal feeding habits by stomach lavage. Marine Mammal Science, 3, 308322.CrossRefGoogle Scholar
Arata, G. F. J. (1954). A note on the flying behavior of certain squids. Nautilus, 68, 13.Google Scholar
Archer, J. (1988). The Behavioural Biology of Aggression. Cambridge: Cambridge University Press.Google Scholar
Aristotle, (1910). Historia Animalium. Translation by D’Arcy Wentworth Thompson. Oxford: Clarendon Press.Google Scholar
Arkhipkin, A. I. & Fedulov, P. P. (1986). Diel movements of juvenile Ilex illecebrosus and other cephalopods in the shelf water–slope water frontal zone off the Scotian Shelf in spring. Journal of Northwest Atlantic Fishery Science, 7, 1524.CrossRefGoogle Scholar
Arnold, J. M. (1962). Mating behavior and social structure in Loligo pealii. Biological Bulletin, 123, 5357.CrossRefGoogle Scholar
Arnold, J. M. (1965). Observations on the mating behavior of the squid Sepioteuthis sepioidea. Bulletin of Marine Science, 15, 216222.Google Scholar
Arnold, J. M. (1984). Cephalopoda. Ch. 6, in The Mollusca, Vol. 7 (ed. Tompa, A. S., Verdonk, N. H. & van de Biggelaar, J. A. M.). Orlando, FL: Academic Press.Google Scholar
Arnold, J. M. (1985). Shell growth, trauma, and repair as an indicator of life history for Nautilus. Veliger, 27, 386396.Google Scholar
Arnold, J. M. (1987). Reproduction and embryology of Nautilus, in Nautilus. The Biology and Paleobiology of a Living Fossil (ed. Saunders, W. B. & Landman, N. H.), pp. 353372. New York: Plenum.CrossRefGoogle Scholar
Arnold, J. M. & Arnold, K. O. (1969). Some aspects of hole-boring predation by Octopus vulgaris. American Zoologist, 9, 991996.CrossRefGoogle Scholar
Arnold, J. M. & Carlson, B. A. (1986). Living Nautilus embryos: preliminary observations. Science, 232, 7376.CrossRefGoogle ScholarPubMed
Arnqvist, G. & Rowe, L. (2013). Sexual Conflict, Monographs in Behavior and Ecology. Princeton, NJ: Princeton University Press.Google Scholar
Aronson, R. B. (1982). An underwater measure of Octopus size. The Veliger, 24, 375377.Google Scholar
Aronson, R. B. (1986). Life history and den ecology of Octopus briareus Robson in a marine lake. Journal of Experimental Marine Biology and Ecology, 95, 3756.CrossRefGoogle Scholar
Aronson, R. B. (1989). The ecology of Octopus briareus Robson in a Bahamian saltwater lake. American Malacological Bulletin, 7, 4756.Google Scholar
Aronson, R. B. (1991). Ecology, paleobiology and evolutionary constraint in the octopus. Bulletin of Marine Science, 49, 245255.Google Scholar
Austin, C. R., Lutwak-Mann, C. & Mann, T. (1964). Spermatophores and spermatozoa of the squid Loligo pealii. Proceedings of the Royal Society of London B, 161, 143152.Google ScholarPubMed
Azuma, A. (1981). Reasoning about flying behavior of squid. Kagaku Asahi, 81–85.Google Scholar
Azuma, A. (2006). The Biokinetics of Flying and Swimming. Reston, VA: American Institute of Aeronautics and Astronautics, Inc.CrossRefGoogle Scholar
Baeg, G. H., Sakurai, Y. & Shimazaki, K. (1993). Maturation processes in female Loligo bleekeri Keferstein (Mollusca: Cephalopoda). Veliger, 36, 228235.Google Scholar
Baglioni, S. (1910). Zur Kenntnis der Leistungen einiger Sinnesorgane (Gesichtssinn, Tastsinn und Geruchssinn) und des Zentralnervensystems der Zephalopoden und Fische. Zeitschrift für Biologie, 53, 255286.Google Scholar
Baker, A. C. (1957). Underwater photographs in the study of oceanic squid. Deep Sea Research, 4, 126138.CrossRefGoogle Scholar
Bakhayokho, M. (1983). Biology of the cuttlefish Sepia officinalis hierredda off the Senegalese coast, in Advances in Assessment of World Cephalopod Resources. FAO Fisheries Technical Paper No. 231 (ed. Caddy, J. F.), pp. 204263: Advances in Assessment of World Cephalopod Resources.Google Scholar
Balda, R. P., Pepperberg, I. M. & Kamil, A. C. (1998). Animal Cognition in Nature: Academic Press.Google Scholar
Banas, P. T., Smith, D. E. & Biggs, D. C. (1982). An association between a pelagic octopod, Argonauta sp. Linnaeus 1758, and aggregate salps. Fishery Bulletin, 80, 648650.Google Scholar
Bandel, K., Reitner, J. & Sturmer, W. (1983). Coleoids from the Lower Devonian Black Slate (‘Hunsruck-Schiefer’) of the Hunsruck (West Germany). Neues Jahrbuch für Geologie und Palaontologie Abhandlungen, 165, 397417.Google Scholar
Barbato, M., Bernard, M., Borrelli, L. & Fiorito, G. (2007). Body patterns in cephalopods: ‘Polyphenism’ as a way of information exchange. Pattern Recognition Letters, 28, 18541864.CrossRefGoogle Scholar
Barber, V. C. & Wright, D. E. (1969). The fine structure of the sense organs of the cephalopod mollusc Nautilus. Zeitschrift für Zellforschung, 102, 293312.CrossRefGoogle ScholarPubMed
Barbosa, A., Allen, J. J., Mäthger, L. M. & Hanlon, R. T. (2012). Cuttlefish use visual cues to determine arm postures for camouflage. Proceedings of the Royal Society B – Biological Sciences, 279, 8490.CrossRefGoogle ScholarPubMed
Barbosa, A., Litman, L. & Hanlon, R. T. (2008). Changeable cuttlefish camouflage is influenced by horizontal and vertical aspects of the visual background. Journal of Comparative Physiology A, 194, 405413.CrossRefGoogle ScholarPubMed
Barbosa, A., Mäthger, L. M., Buresch, K. C. et al. (2008). Cuttlefish camouflage: the effects of substrate contrast and size in evoking uniform, mottle or disruptive body patterns. Vision Research, 48, 12421253.CrossRefGoogle ScholarPubMed
Barbosa, A., Mäthger, L. M., Chubb, C. et al. (2007). Disruptive coloration in cuttlefish: a visual perception mechanism that regulates ontogenetic adjustment of skin patterning. Journal of Experimental Biology, 210, 11391147.CrossRefGoogle ScholarPubMed
Barnes, R. S. K. (1991). Reproduction, life histories and dispersal, in Fundamentals of Aquatic Ecology (ed. Barnes, R. S. K. & Mann, K. H.), pp. 145171. Oxford: Blackwell Scientific Publications.CrossRefGoogle Scholar
Barrows, E. M. (1992). The Complete Animal Behavior Desk Reference. Boca Raton, Florida: CRC Press.Google Scholar
Bartol, I. K., Mann, R. & Vecchione, M. (2002). Distribution of the euryhaline squid Lolliguncula brevis in Chesapeake Bay: effects of selected abiotic factors. Marine Ecology Progress Series, 226, 235247.CrossRefGoogle Scholar
Bas, C. (1979). Un modelo de distribución de dos especies: Pagellus acarne y Octopus vulgaris, influidas por la pesca y las condiciónes ambientales. Investigación Pesquera, 43, 141148.Google Scholar
Basil, J. A., Bahctinova, I., Kuroiwa, K. et al. (2005). The function of the rhinophore and the tentacles of Nautilus pompilius L. (Cephalopoda, Nautiloidea) in orientation to odor. Marine and Freshwater Behaviour and Physiology, 38, 209221.CrossRefGoogle Scholar
Basil, J. A., Barord, G., Crook, R. J. et al. (2011). A synthetic approach to the study of learning and memory in Chambered Nautilus L. (Cephalopoda, Nautiloidea). Vie et Milieu – Life and Environment, 61, 231242.Google Scholar
Basil, J. A., Hanlon, R. T., Sheikh, S. I. & Atema, J. (2000). Three-dimensional odor tracking by Nautilus pompilius. Journal of Experimental Biology, 203, 14091414.CrossRefGoogle ScholarPubMed
Basil, J. A., Lazenby, G. B., Nakanuku, L. & Hanlon, R. T. (2002). Female Nautilus are attracted to male conspecific odor. Bulletin of Marine Science, 70, 217225.Google Scholar
Bateman, A. W., Vos, M. & Anholt, B. R. (2014). When to defend: antipredator defenses and the predation sequence. The American Naturalist, 183, 847855.CrossRefGoogle ScholarPubMed
Bateson, P. P. G. (1991). Levels and processes, in The Development and Integration of Behaviour (ed. Bateson, P. P. G.), pp. 316. Cambridge: Cambridge University Press.Google Scholar
Bazzino, G., Gilly, W. F., Markaida, U., Salinas-Zavala, C. A. & Ramos-Castillejos, J. (2010). Horizontal movements, vertical-habitat utilization and diet of the jumbo squid (Dosidicus gigas) in the Pacific Ocean off Baja California Sur, Mexico. Progress in Oceanography, 86, 5971.CrossRefGoogle Scholar
Beja, P. R. (1991). Diet of otters (Lutra lutra) in closely associated freshwater, brackish and marine habitats in south-west Portugal. Journal of Zoology (London), 225, 141152.CrossRefGoogle Scholar
Bell, G. R. R., Kuzirian, A. M., Senft, S. L. et al. (2013). Chromatophore radial muscle fibers anchor in flexible squid skin. Invertebrate Biology, 132, 120132.CrossRefGoogle Scholar
Bell, G. R. R., Mäthger, L. M., Gao, M. et al. (2014). Diffuse white structural coloration from multilayer reflectors in a squid. Advanced Materials, 26, 43524356.CrossRefGoogle Scholar
Bello, G. (1991). Role of cephalopods in the diet of the swordfish, Xiphias gladius, from the eastern Mediterranean Sea. Bulletin of Marine Science, 49, 312324.Google Scholar
Benchley, P. (1991). Beast. New York: Random House.Google Scholar
Benoit-Bird, K. J. & Gilly, W. F. (2012). Coordinated nocturnal behavior of foraging jumbo squid Dosidicus gigas. Marine Ecology Progress Series, 455, 211228.CrossRefGoogle Scholar
Benoit-Bird, K. J., Gilly, W. F., Au, W. W. L. & Mate, B. (2008). Controlled and in situ target strengths of the jumbo squid Dosidicus gigas and identification of potential acoustic scattering sources. Journal of the Acoustical Society of America, 123, 13181328.CrossRefGoogle ScholarPubMed
Bergmann, S., Lieb, B., Ruth, P. & Markl, J. (2006). The hemocyanin from a living fossil, the cephalopod Nautilus pompilius: protein structure, gene organization, and evolution. Journal of Molecular Evolution, 62, 362374.Google Scholar
Bergstrom, B. & Summers, W. C. (1983). Sepietta oweniana, in Cephalopod Life Cycles, Vol. I: Species Accounts (ed. Boyle, P. R.), pp. 7591. London: Academic Press.Google Scholar
Berruti, A. & Harcus, T. (1978). Cephalopod prey of the Sooty Albatrosses Phoebetri fusca and P. palpebrata at Marion Island. South African Journal of Antarctic Research, 8, 99103.Google Scholar
Bert, P. (1867). Mémoire sur la physiologie de la seiche (Sepia officinalis, Linn.). Mémoires de la Societe des Sciences Physiques et Naturelles de Bordeaux, 5, 115138.Google Scholar
Biagi, V. & Bello, G. (2009). Occurrence of an egg mass of Thysanoteuthis rhombus (Cephalopoda: Teuthida) in the Strait of Messina (Italy), locus typicus of the species. Bollettino Malacologia, 45, 3538.Google Scholar
Bidder, A. M. (1962). Use of the tentacles, swimming, and buoyancy control in the pearly nautilus. Nature, 196, 451454.CrossRefGoogle Scholar
Bidder, A. M. (1966). Feeding and digestion in cephalopods, in Physiology of Mollusca, Vol. 2 (ed. Wilbur, K. M. & Yonge, C. M.), pp. 97124. New York: Academic Press.CrossRefGoogle Scholar
Biedermann, G. B. & Davey, V. A. (1993). Social-learning in invertebrates. Science, 259, 16271628.CrossRefGoogle Scholar
Birkhead, T. R. (1989). The intelligent sperm? A concise review of sperm competition. Journal of Zoology (London), 218, 347351.CrossRefGoogle Scholar
Birkhead, T. R. & Moller, A. P. (1998). Sperm Competition and Sexual Selection. New York: Academic Press.Google Scholar
Birkhead, T. R. & Parker, G. A. (1997). Sperm competition and mating systems, in Behavioural Ecology: An Evolutionary Approach (ed. Krebs, J. R. & Davies, N. B.), pp. 121145. Oxford: Blackwell Science, Ltd.Google Scholar
Birkhead, T. R. & Pizzari, T. (2002). Postcopulatory sexual selection. Nature Reviews Genetics, 3, 262273.CrossRefGoogle ScholarPubMed
Birkhead, T. R., Hosken, D. J. & Pitnick, S. S. (2009). Sperm Biology: An Evolutionary Perspective: Academic Press.Google Scholar
Bitterman, M. E. (1975). Critical commentary, in Invertebrate Learning (ed. Corning, W. C., Dyal, J. A. & Willows, A. O. D.), pp. 139145. New York: Plenum.Google Scholar
Bleckmann, H., Budelmann, B. U. & Bullock, T. H. (1991). Peripheral and central nervous responses evoked by small water movements in a cephalopod. Journal of Comparative Physiology A, 168, 247257.Google Scholar
Boal, J. G. (1991). Complex learning in Octopus bimaculoides. American Malacological Bulletin, 9, 7580.Google Scholar
Boal, J. G. (1996). Absence of social recognition in laboratory-reared cuttlefish, Sepia officinalis L. (Mollusca: Cephalopoda). Animal Behaviour, 52, 529537.Google Scholar
Boal, J. G. (1997). Female choice of males in cuttlefish (Mollusca: Cephalopoda). Behaviour, 134, 975988.CrossRefGoogle Scholar
Boal, J. G. (2006). Social recognition: a top down view of cephalopod behavior. Vie et Milieu – Life and Environment, 56, 6979.Google Scholar
Boal, J. G. (2011). Behavioral research methods for octopuses and cuttlefishes. Vie et Milieu – Life and Environment, 61, 203210.Google Scholar
Boal, J. G. & Fenwick, J. W. (2007). Laterality in octopus eye use? Animal Behaviour, 74, E1E2.CrossRefGoogle Scholar
Boal, J. G. & Golden, D. K. (1999). Distance chemoreception in the common cuttlefish, Sepia officinalis (Mollusca, Cephalopoda). Journal of Experimental Marine Biology and Ecology, 235, 307317.CrossRefGoogle Scholar
Boal, J. G. & Gonzalez, S. A. (1998). Social behaviour of individual oval squids (Cephalopoda, Teuthoidea, Loliginidae, Sepioteuthis lessoniana) within a captive school. Ethology, 104, 161178.CrossRefGoogle Scholar
Boal, J. G. & Marsh, S. E. (1998). Social recognition using chemical cues in cuttlefish (Sepia officinalis Linnaeus, 1758). Journal of Experimental Marine Biology and Ecology, 230, 183192.CrossRefGoogle Scholar
Boal, J. G., Dunham, A. W., Williams, K. T. & Hanlon, R. T. (2000). Experimental evidence for spatial learning in octopuses (Octopus bimaculoides). Journal of Comparative Psychology, 114, 246252.CrossRefGoogle ScholarPubMed
Boal, J. G., Hylton, R. A., Gonzalez, S. A. & Hanlon, R. T. (1999). Effects of crowding on the social behavior of cuttlefish (Sepia officinalis). Contemporary Topics in Laboratory Animal Science, 38, 4955.Google Scholar
Boal, J. G., Prosser, K. N., Holm, J. B. et al. (2010). Sexually mature cuttlefish are attracted to the eggs of conspecifics. Journal of Chemical Ecology, 36, 834836.CrossRefGoogle Scholar
Boal, J. G., Shashar, N., Grable, M. M. et al. (2004). Behavioral evidence for intraspecific signals with achromatic and polarized light by cuttlefish (Mollusca: Cephalopoda). Behaviour, 141, 837861.Google Scholar
Boal, J. G., Wittenberg, K. M. & Hanlon, R. T. (2000). Observational learning does not explain improvement in predation tactics by cuttlefish (Mollusca: Cephalopoda). Behavioural Processes, 52, 141153.Google Scholar
Boletzky, S. von (1977). Post-hatching behaviour and mode of life in cephalopods. Symposia of the Zoological Society of London, 38, 557567.Google Scholar
Boletzky, S. von (1978). Nos connaissances actuelles sur le développement des octopodes. Vie et Milieu, 28 /29, 85120.Google Scholar
Boletzky, S. von (1981). Réflexions sur les stratégies de reproduction chez les céphalopodes. Extrait du Bulletin de la Société Zoologique de France, 106, 293304.Google Scholar
Boletzky, S. von (1983a). Sepiola robusta, in Cephalopod Life Cycles, Vol. 1. Species Accounts (ed. Boyle, P. R.), pp. 5367. London: Academic Press.Google Scholar
Boletzky, S. von (1983b). Sepia officinalis, in Cephalopod Life Cycles, Vol. 1: Species Accounts (ed. Boyle, P. R.), pp. 3152. London: Academic Press.Google Scholar
Boletzky, S. von (1987a). Juvenile behaviour, in Cephalopod Life Cycles, Vol. 2: Comparative Reviews (ed. Boyle, P. R.), pp. 4560. London: Academic Press.Google Scholar
Boletzky, S. von (1987b). Fecundity variation in relation to intermittent or chronic spawning in the cuttlefish, Sepia officinalis L. (Mollusca, Cephalopoda). Bulletin of Marine Science, 40, 382387.Google Scholar
Boletzky, S. von (1988). A new record of long-continued spawning in Sepia officinalis (Mollusca, Cephalopoda). Rapport Commission Internationale pour la Mer Mediterranee, 31, 257.Google Scholar
Boletzky, S. von & Boletzky, M. V. von (1970). Das Eingraben in Sand bei Sepiola und Sepietta (Mollusca, Cephalopoda). Revue Suisse de Zoologie, 77, 536548.CrossRefGoogle Scholar
Boletzky, S. von & Hanlon, R. T. (1983). A review of the laboratory maintenance, rearing and culture of cephalopod molluscs. Memoirs of the National Museum of Victoria, 44, 147187.CrossRefGoogle Scholar
Boletzky, S. von, Boletzky, M. V. von, Frosch, D. & Gatzi, V. (1971). Laboratory rearing of Sepiolinae (Mollusca: Cephalopoda). Marine Biology, 8, 8287.Google Scholar
Boletzky, S. von, Rio, M. & Roux, M. (1992). Octopod ‘ballooning’ response. Nature, 356, 199.Google Scholar
Bond, A. B. (1989a). Toward a resolution of the paradox of aggressive displays: I. Optimal deceit in the communication of fighting ability. Ethology, 81, 2946.CrossRefGoogle Scholar
Bond, A. B. (1989b). Toward a resolution of the paradox of aggressive displays: II. Behavioral efference and the communication of intentions. Ethology, 81, 235249.Google Scholar
Bone, Q. & Marshall, N. B. (1982). Biology of Fishes. Glasgow: Blackie.Google Scholar
Bone, Q., Brown, E. R. & Travers, G. (1994). On the respiratory flow in the cuttlefish Sepia officinalis. Journal of Experimental Biology, 194, 153165.Google Scholar
Bone, Q., Pulsford, A. & Chubb, A. D. (1981). Squid mantle muscle. Journal of the Marine Biological Association of the United Kingdom, 61, 327342.CrossRefGoogle Scholar
Borrelli, L., Gherardi, F. & Fiorito, G. (2006). A Catalog of Body Patterning in Cephalopoda. Firenze: Firenze University Press.CrossRefGoogle Scholar
Bott, R. (1938). Kopula und Eiablage von Sepia officinalis L. Zeitschrift für Morphologie und Oekologie der Tiere, 34, 150160.Google Scholar
Boucaud-Camou, E. & Boismery, J. (1991). The migrations of the cuttlefish (Sepia officinalis L) in the English Channel, in First International Symposium on the Cuttlefish Sepia (ed. Boucaud-Camou, E.), pp. 1358. Caen: Centre de Publications de l’Universite de Caen.Google Scholar
Boucaud-Camou, E. & Boucher-Rodoni, R. (1983). Feeding and digestion in cephalopods, in The Mollusca, Vol. 5: Physiology, Part 2 (ed. Saleuddin, A. S. M. & Wilbur, K. M.), pp. 149187. New York: Academic Press.CrossRefGoogle Scholar
Boucher-Rodoni, R., Boucaud-Camou, E. & Mangold, K. (1987). Feeding and digestion, in Cephalopod Life Cycles, Vol. 2: Comparative Reviews (ed. Boyle, P. R.), pp. 85108. London: Academic Press.Google Scholar
Bouligand, Y. (1961). Le dispositif d’accrochage des oeufs de Sepia elegans sur Alcyonium palmatum. Vie et Milieu, 12, 589594.Google Scholar
Bouth, H. F., Leite, T. S., de Lima, F. D. & Oliveira, J. E. L. (2011). Atol das Rocas: an oasis for Octopus insularis juveniles (Cephalopoda: Octopodidae). Zoologia, 28, 4552.CrossRefGoogle Scholar
Boutilier, R. G., West, T. G., Pogson, G. H. et al. (1996). Nautilus and the art of metabolic maintenance. Nature, 382, 534536.CrossRefGoogle Scholar
Boutilier, R. G., West, T. G., Webber, D. M. et al. (2000). The protective effects of hypoxia-induced hypometabolism in the Nautilus. Journal of Comparative Physiology B – Biochemical Systemic and Environmental Physiology, 170, 261268.CrossRefGoogle ScholarPubMed
Bouwma, P. E. & Herrnkind, W. F. (2009). Sound production in Caribbean spiny lobster Panulirus argus and its role in escape during predatory attack by Octopus briareus. New Zealand Journal of Marine and Freshwater Research, 43, 313.CrossRefGoogle Scholar
Bower, J. R. & Ichii, T. (2005). The red flying squid (Ommastrephes bartramii): a review of recent research and the fishery in Japan. Fisheries Research, 76, 3955.CrossRefGoogle Scholar
Bower, J. R. & Sakurai, Y. (1996). Laboratory observations on Todarodes pacificus (Cephalopoda: Ommastrephidae) egg masses. American Malacological Bulletin, 13, 6571.Google Scholar
Bower, J. R., Seki, K., Kubodera, T., Yamamoto, J. & Nobetsu, T. (2012). Brooding in a gonatid squid off northern Japan. Biological Bulletin, 223, 259262.CrossRefGoogle Scholar
Bowman, R. E., Stillwell, C. E., Michaels, W. L. & Grosslein, M. D. (2000). Food of northwest Atlantic fishes and two common species of squid. NOAA Technical Memorandum NMFS-NE-155, 1–135.Google Scholar
Boycott, B. B. (1953). The chromatophore system of cephalopods. Proceedings of the Linnaean Society (London), 164, 235240.Google Scholar
Boycott, B. B. (1954). Learning in Octopus vulgaris and other cephalopods. Pubblicazioni della Stazione Zoologica di Napoli, 25, 6793.Google Scholar
Boycott, B. B. (1958). The cuttlefish – Sepia. New Biology, 25, 98118.Google Scholar
Boycott, B. B. (1960). The functioning of the statocysts of Octopus vulgaris. Proceedings of the Royal Society of London B, 152, 7887.Google Scholar
Boycott, B. B. (1961). The functional organization of the brain of the cuttlefish Sepia officinalis. Proceedings of the Royal Society of London B, 153, 503534.Google Scholar
Boycott, B. B. (1965). A comparison of living Sepioteuthis sepioidea and Doryteuthis plei with other squids, and with Sepia officinalis. Journal of Zoology (London), 147, 344351.Google Scholar
Boycott, B. B. & Young, J. Z. (1950). The comparative study of learning. Symposia of the Society for Experimental Biology, 4, 432453.Google Scholar
Boycott, B. B. & Young, J. Z. (1955a). A memory system in Octopus vulgaris Lamarck. Proceedings of the Royal Society B – Biological Sciences, 143, 449.Google ScholarPubMed
Boycott, B. B. & Young, J. Z. (1955b). Memories controlling attacks on food objects by Octopus vulgaris Lamarck. Pubblicazioni della Stazione Zoologica di Napoli, 27, 232249.Google Scholar
Boycott, B. B. & Young, J. Z. (1956). The subpedunculate body and nerve and other organs associated with the optic tract of cephalopods, in Bertil Hanström. Zoological Papers in Honour of his Sixty-Fifth Birthday (ed. Wingstrand, K. G.), pp. 76105. Lund, Sweden: Zoological Institute.Google Scholar
Boycott, B. B. & Young, J. Z. (1957). Effects of interference with the vertical lobe on visual discrimination in Octopus vulgaris Lamarck. Proceedings of the Royal Society of London B, 146, 439459.Google Scholar
Boyd, I. L. & Arnbom, T. (1991). Diving behaviour in relation to water temperature in the southern elephant seal: foraging implications. Polar Biology, 11, 259266.Google Scholar
Boyle, P. R. (1980). Home occupancy by male Octopus vulgaris in a large seawater tank. Animal Behaviour, 28, 11231126.Google Scholar
Boyle, P. R. (1983a). Ventilation rate and arousal in the octopus. Journal of Experimental Marine Biology and Ecology, 69, 129136.CrossRefGoogle Scholar
Boyle, P. R. (1983b). Cephalopod Life Cycles, Vol. 1: Species Accounts. London: Academic Press.Google Scholar
Boyle, P. R. (1986a). Responses to water-borne chemicals by the octopus Eledone cirrhosa (Lamarck, 1798). Journal of Experimental Marine Biology and Ecology, 104, 2330.Google Scholar
Boyle, P. R. (1986b). Neural control of cephalopod behavior, in The Mollusca, Vol. 9: Neurobiology and Behavior (ed. Willows, A. O. D.), pp. 199. Orlando: Academic Press.Google Scholar
Boyle, P. R. (1986c). Report on a specimen of Architeuthis stranded near Aberdeen, Scotland. Journal of Molluscan Studies, 52, 8182.Google Scholar
Boyle, P. R. (1987). Cephalopod Life Cycles, Vol. 2: Comparative Reviews. London: Academic Press.Google Scholar
Boyle, P. R. (1990). Prey handling and salivary secretions in octopuses, in Trophic Relationships in the Marine Environment (ed. Barnes, M. & Gibson, R. N.), pp. 541552. Aberdeen: University Press.Google Scholar
Boyle, P. R. & Dubas, F. (1981). Components of body pattern displays in the octopus Eledone cirrhosa (Mollusca: Cephalopoda). Marine Behaviour and Physiology, 8, 135148.Google Scholar
Boyle, P. R. & Knobloch, D. (1981). Hole boring of crustacean prey by the octopus Eledone cirrhosa (Mollusca, Cephalopoda). Journal of Zoology (London), 193, 110.Google Scholar
Boyle, P. R. & Rodhouse, P. (2005). Cephalopods. Ecology and Fisheries: Oxford: Blackwell Science.CrossRefGoogle Scholar
Boyle, P. R., Pierce, G. J. & Hastie, L. C. (1995). Flexible reproductive strategies in the squid Loligo forbesi. Marine Biology, 121, 501508.Google Scholar
Bradbury, J. W. & Vehrencamp, S. L. (2011). Principles of Animal Communication. Sunderland, MA: Sinauer Associates, Inc.Google Scholar
Bradley, E. A. & Messenger, J. B. (1977). Brightness preference in Octopus as a function of the background brightness. Marine Behaviour and Physiology, 4, 243251.CrossRefGoogle Scholar
Bradley, E. A. & Young, J. Z. (1975). Comparison of visual and tactile learning in Octopus after lesions to one of two memory systems. Journal of Neuroscience Research, 1, 185205.Google Scholar
Braid, H. E. & Bolstad, K. S. R. (2014). Feeding ecology of the largest mastigoteuthid squid species, Idioteuthis cordiformis (Cephalopoda, Mastigoteuthidae). Marine Ecology Progress Series, 515, 275279.CrossRefGoogle Scholar
Braithwaite, V. (2010). Do Fish Feel Pain? Oxford: Oxford University Press.Google Scholar
Briffa, M. (2013). Plastic proteans: reduced predictability in the face of predation risk in hermit crabs. Biology Letters, 9, 20130592.CrossRefGoogle ScholarPubMed
Briffa, M. & Hardy, I. C. W. (2013). Introduction to animal contests, in Animal Aggression (ed. Hardy, I. C. W. & Briffa, M.), pp. 14. Cambridge: Cambridge University Press.Google Scholar
Brocco, S. L. (1971). Aspects of the Biology of the Sepiolid Squid Rossia pacifica Berry. Victoria, British Columbia: University of Victoria.Google Scholar
Brooks, W. R. (1988). The influence of the location and abundance of the sea anemone Calliactis tricolor (Le Sueur) in protecting hermit crabs from octopus predators. Journal of Experimental Marine Biology and Ecology, 116, 1521.CrossRefGoogle Scholar
Brown, C., Garwood, M. P. & Williamson, J. E. (2012). It pays to cheat: tactical deception in a cephalopod social signalling system. Biology Letters, 8, 729732.Google Scholar
Brown, E. R., Piscopo, S., De Stefano, R. & Giuditta, A. (2006). Brain and behavioural evidence for rest–activity cycles in Octopus vulgaris. Behavioural Brain Research, 172, 355359.CrossRefGoogle ScholarPubMed
Brown, P. K. & Brown, P. S. (1958). Visual pigments of the octopus and cuttlefish. Nature, 182, 12881290.Google Scholar
Bruun, A. F. (1943). The biology of Spirula spirula (L). Dana Reports, 24, 144.Google Scholar
Budelmann, B. U. (1970). Die Arbeitsweise de Statolithenorgane von Octopus vulgaris. Zeitschrift für vergleichende Physiologie, 70, 278312.Google Scholar
Budelmann, B. U. (1976). Equilibrium receptor systems in molluscs, in Structure and Function of Proprioceptors in the Invertebrates (ed. Mill, P. J.), pp. 529566. London: Chapman & Hall.Google Scholar
Budelmann, B. U. (1990). The statocysts of squid, in Squid as Experimental Animals (ed. Gilbert, D. L., Adelman, W. J. & Arnold, J. M.), pp. 421439. New York: Plenum Press.Google Scholar
Budelmann, B. U. (1992). Hearing in non-arthropod invertebrates, in The Evolution of Hearing (ed. Webster, D. B., Fay, R. R. & Popper, A. N.), pp. 141155. New York: Springer.CrossRefGoogle Scholar
Budelmann, B. U. (1994). Cephalopod sense organs, nerves and the brain: adaptations for high performance and life style. Marine and Freshwater Behaviour and Physiology, 25, 1333.Google Scholar
Budelmann, B. U. (1995). The cephalopod nervous system: what evolution has made of the molluscan design, in The Nervous Systems of Invertebrates. An Evolutionary and Comparative Approach (ed. Breidbach, O. & Kutsch, W.), pp. 115138. Basel: Birkhauser.CrossRefGoogle Scholar
Budelmann, B. U. (1996). Active marine predators: the sensory world of cephalopods. Marine and Freshwater Behaviour and Physiology, 27, 5975.CrossRefGoogle Scholar
Budelmann, B. U. & Bleckmann, H. (1988). A lateral line analogue in cephalopods: water waves generate microphonic potentials in the epidermal head lines of Sepia and Lolliguncula. Journal of Comparative Physiology A, 164, 15.CrossRefGoogle ScholarPubMed
Budelmann, B. U. & Young, J. Z. (1984). The statocyst–oculomotor system of Octopus vulgaris: extraocular eye muscles, eye muscle nerves, statocyst nerves and the oculomotor centre in the central nervous system. Philosophical Transactions of the Royal Society of London B, 306, 159189.Google Scholar
Budelmann, B. U. & Young, J. Z. (1993). The oculomotor system of decapod cephalopods: eye muscles, eye muscle nerves, and the oculomotor neurons in the central nervous system. Philosophical Transactions of the Royal Society of London B, 340, 93125.Google Scholar
Budelmann, B. U., Bullock, T. H. & Williamson, R. (1995). Cephalopod brains: promising preparations for brain physiology, in Cephalopod Neurobiology (ed. Abbott, N. J., Williamson, R. & Maddock, L.). Oxford: Oxford University Press.Google Scholar
Budelmann, B. U., Riese, U. & Bleckmann, H. (1991). Structure, function, biological significance of the cuttlefish ‘lateral lines’, in The Cuttlefish. First International Symposium on the Cuttlefish Sepia (ed. Boucaud-Camou, E.), pp. 201209. Caen: Centre de Publications de l’Universite de Caen.Google Scholar
Budelmann, B. U., Sachse, M. & Staudigl, M. (1987). The angular acceleration receptor system of Octopus vulgaris: morphometry, ultrastructure, and neuronal and synaptic organization. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 315, 305343.Google Scholar
Budelmann, B. U., Schipp, R. & Boletzky, S. von (1997). Cephalopoda, in Microscopic Anatomy of Invertebrates, Vol. 6A: Mollusca II, pp. 119414. New York: Wiley-Liss, Inc.Google Scholar
Bullock, T. H. (1965). Mollusca: Cephalopoda, in Structure and Function in the Nervous Systems of Invertebrates (ed. Bullock, T. H. & Horridge, G. A.), pp. 14331515. San Francisco: Freeman.Google Scholar
Bullock, T. H. & Basar, E. (1988). Comparison of ongoing compound field potentials in the brains of invertebrates and vertebrates. Brain Research Reviews, 13, 5775.Google Scholar
Bullock, T. H. & Budelmann, B. U. (1991). Sensory evoked potentials in unanesthetized unrestrained cuttlefish: a new preparation for brain physiology in cephalopods. Journal of Comparative Physiology A, 168, 141150.Google Scholar
Buresch, K. C., Boal, J. G., Knowles, J. et al. (2003). Contact chemosensory cues in egg bundles elicit male–male agonistic conflicts in the squid Loligo pealeii. Journal of Chemical Ecology, 29, 547560.CrossRefGoogle ScholarPubMed
Buresch, K. C., Boal, J. G., Nagle, G. T. et al. (2004). Experimental evidence that ovary and oviducal gland extracts influence male agonistic behavior in squids. Biological Bulletin, 206, 13.Google Scholar
Buresch, K. C., Gerlach, G. & Hanlon, R. T. (2006). Multiple genetic stocks of the longfin inshore squid Loligo pealeii in the NW Atlantic: stocks segregate inshore in summer, but aggregate offshore in winter. Marine Ecology Progress Series, 310, 263270.Google Scholar
Buresch, K. C., Mäthger, L. M., Allen, J. J. et al. (2011). The use of background matching vs. masquerade for camouflage in cuttlefish Sepia officinalis. Vision Research, 51, 23622368.CrossRefGoogle ScholarPubMed
Buresch, K. C., Maxwell, M. R., Cox, M. R. & Hanlon, R. T. (2009). Temporal dynamics of mating and paternity in the squid Loligo pealeii. Marine Ecology Progress Series, 387, 197203.Google Scholar
Buresch, K. M., Hanlon, R. T., Maxwell, M. R. & Ring, S. (2001). Microsatellite DNA markers indicate a high frequency of multiple paternity within individual field-collected egg capsules of the squid Loligo pealeii. Marine Ecology Progress Series, 210, 161165.CrossRefGoogle Scholar
Burghardt, G. M. (1999). Conceptions of play and the evolution of animal minds. Evolutionary Cognition, 5, 115123.Google Scholar
Burghardt, G. M. (2005). The Genesis of Animal Play. Cambridge, MA: MIT Press.CrossRefGoogle Scholar
Bush, S. L. (2012). Economy of arm autotomy in the mesopelagic squid Octopoteuthis deletron. Marine Ecology Progress Series, 458, 133140.CrossRefGoogle Scholar
Bush, S. L. & Robison, B. H. (2007). Ink utilization by mesopelagic squid. Marine Biology, 152, 485494.Google Scholar
Bush, S. L., Hoving, H. J. T., Huffard, C. L., Robison, B. H. & Zeidberg, L. D. (2012). Brooding and sperm storage by the deep-sea squid Bathyteuthis berryi (Cephalopoda: Decapodiformes). Journal of the Marine Biological Association of the United Kingdom, 92, 16291636.Google Scholar
Bush, S. L., Robison, B. H. & Caldwell, R. L. (2009). Behaving in the dark: locomotor, chromatic, postural, and bioluminescent behaviors of the deep-sea squid Octopoteuthis deletron Young 1972. Biological Bulletin, Marine Biological Laboratory, Woods Hole, 216, 722.Google Scholar
Butler, M. J. & Lear, J. A. (2009). Habitat-based intraguild predation by Caribbean reef octopus Octopus briareus on juvenile Caribbean spiny lobster Panulirus argus. Marine Ecology Progress Series, 386, 115122.CrossRefGoogle Scholar
Butterworth, M. (1982). Shell utilization by Octopus joubini. Tallahassee, FL: Unpublished M.S. Thesis, Florida State University.Google Scholar
Byrne, R. A., Kuba, M. & Griebel, U. (2002). Lateral asymmetry of eye use in Octopus vulgaris. Animal Behaviour, 64, 461468.CrossRefGoogle Scholar
Byrne, R. A., Kuba, M. J. & Meisel, D. V. (2004). Lateralized eye use in Octopus vulgaris shows antisymmetrical distribution. Animal Behaviour, 68, 11071114.CrossRefGoogle Scholar
Byrne, R. A., Kuba, M. J., Meisel, D. V., Griebel, U. & Mather, J. A. (2006a). Does Octopus vulgaris have preferred arms? Journal of Comparative Psychology, 120, 198204.CrossRefGoogle ScholarPubMed
Byrne, R. A., Kuba, M. J., Meisel, D. V., Griebel, U. & Mather, J. A. (2006b). Octopus arm choice is strongly influenced by eye use. Behavioural Brain Research, 172, 195201.CrossRefGoogle ScholarPubMed
Byrne, R. W. (2000). Animal cognition in nature. Trends in Cognitive Sciences, 4, 73.CrossRefGoogle Scholar
Byrne, R. W. & Bates, L. A. (2006). Why are animals cognitive? Current Biology, 16, R445R448.Google Scholar
Byzov, A. L., Orlov, O. Y. & Utina, I. A. (1962). An investigation of adaptation in the eyes of cephalopod molluscs. Biofizika, 7, 318327.Google Scholar
Caddy, J. F. (1983). Advances in Assessment of World Cephalopod Resources. FAO, United Nations, Rome: FAO Fisheries Technical Paper No. 231.Google Scholar
Caldwell, R. L. (2005). An observation of inking behavior protecting adult Octopus bocki from predation by green turtle (Chelonia mydas) hatchlings. Pacific Science, 59, 6972.Google Scholar
Caldwell, R. L. & Lamp, K. (1981). Chemically mediated recognition by the stomatopod Gonodactylus bredini of its competitor, the octopus Octopus joubini. Marine Behaviour and Physiology, 8, 3541.Google Scholar
Calisti, M., Giorelli, M., Levy, G. et al. (2011). An octopus-bioinspired solution to movement and manipulation for soft robots. Bioinspiration and Biomimetics, 6, 036002.Google Scholar
Callan, H. G. (1940). Absence of a sex-hormone controlling regeneration of the hectocotylus in Octopus vulgaris Lam. Pubblicazioni della Stazione Zoologica di Napoli, 18, 1519.Google Scholar
Calow, P. (1987). Fact and theory – an overview, in Cephalopod Life Cycles. Vol. 2: Comparative Reviews (ed. Boyle, P. R.), pp. 351365. London: Academic Press.Google Scholar
Carere, C., Wood, J. B. & Mather, J. (2011). Species differences in captivity: where are the invertebrates? Trends in Ecology and Evolution, 26, 211211.Google Scholar
Cariello, L. & Zanetti, L. (1977). Alpha- and beta-cephalotoxin: two paralysing proteins from posterior salivary glands of Octopus vulgaris. Comparative Biochemistry and Physiology, C57, 169173.Google Scholar
Carlson, B. A. (1987). Collection and aquarium maintenance of Nautilus, in Nautilus. The Biology and Paleobiology of a Living Fossil (ed. Saunders, W. B. & Landman, N. H.), pp. 563578. New York: Plenum.Google Scholar
Carlson, B. A., Awai, M. L. & Arnold, J. M. (1992). Hatching and early growth of Nautilus belauensis and implications on the distribution of the genus Nautilus. Proceedings of the Seventh International Coral Reef Symposium, Guam, 1, 587592.Google Scholar
Carlson, B. A., McKibben, J. N. & DeGruy, M. V. (1984). Telemetric investigation of vertical migration of Nautilus belauensis in Palau. Pacific Science, 38, 183188.Google Scholar
Caro, T. (2005). Antipredator Defenses in Birds and Mammals. Chicago: University of Chicago Press.Google Scholar
Cartron, L., Darmaillacq, A. S., Jozet-Alves, C., Shashar, N. & Dickel, L. (2012). Cuttlefish rely on both polarized light and landmarks for orientation. Animal Cognition, 15, 591596.CrossRefGoogle ScholarPubMed
Cartron, L., Shashar, N., Dickel, L. & Darmaillacq, A. S. (2013). Effects of stimuli shape and polarization in evoking deimatic patterns in the European cuttlefish, Sepia officinalis, under varying turbidity conditions. Invertebrate Neuroscience, 13, 1926.Google Scholar
Catchpole, C. K. (1980). Sexual selection and the evolution of complex songs among European warblers of the genus Acrocephalus. Behaviour, 74, 149166.Google Scholar
Chamberlain, J. A. (1987). Locomotion of Nautilus, in Nautilus. The Biology and Paleobiology of a Living Fossil (ed. Saunders, W. B. & Landman, N. H.), pp. 489525. New York: Plenum.Google Scholar
Chance, M. R. A. & Russell, W. M. S. (1959). Protean displays: a form of allaesthetic behaviour. Proceedings of the Zoological Society of London, 132, 6570.CrossRefGoogle Scholar
Chapman, T. (2006). Evolutionary conflicts of interest between males and females. Current Biology, 16, R744R754.CrossRefGoogle ScholarPubMed
Chase, R. & Wells, M. J. (1986). Chemotactic behaviour in Octopus. Journal of Comparative Physiology A, 158, 375381.CrossRefGoogle Scholar
Chedekel, M. R., Murr, B. L. & Zeise, L. (1992). Melanin standard method: empirical formula. Pigment Cell Research, 5, 143147.CrossRefGoogle ScholarPubMed
Cheng, M. W. & Caldwell, R. L. (2000). Sex identification and mating in the blue-ringed octopus, Hapalochlaena lunulata. Animal Behaviour, 60, 2733.CrossRefGoogle ScholarPubMed
Cherel, Y. & Hobson, K. A. (2005). Stable isotopes, beaks and predators: a new tool to study the trophic ecology of cephalopods, including giant and colossal squids. Proceedings of the Royal Society B – Biological Sciences, 272, 16011607.Google Scholar
Chiao, C. C. & Hanlon, R. T. (2001a). Cuttlefish camouflage: visual perception of size, contrast and number of white squares on artificial chequerboard substrata initiates disruptive coloration. Journal of Experimental Biology, 204, 21192125.Google Scholar
Chiao, C. C. & Hanlon, R. T. (2001b). Cuttlefish cue visually on area – not shape or aspect ratio – of light objects in the substrate to produce disruptive body patterns for camouflage. Biological Bulletin, 201, 269270.Google Scholar
Chiao, C. C., Chubb, C., Buresch, K. C. et al. (2010). Mottle camouflage patterns in cuttlefish: quantitative characterization and visual background stimuli that evoke them. Journal of Experimental Biology, 213, 187199.Google Scholar
Chiao, C. C., Chubb, C. & Hanlon, R. T. (2007). Interactive effects of size, contrast, intensity and configuration of background objects in evoking disruptive camouflage in cuttlefish. Vision Research, 47, 22232235.Google Scholar
Chiao, C. C., Kelman, E. J. & Hanlon, R. T. (2005). Disruptive body pattern of cuttlefish (Sepia officinalis) requires visual information regarding edges and contrast of objects in natural substrate backgrounds. Biological Bulletin, 208, 711.Google Scholar
Chiao, C. C., Ulmer, K. M., Siemann, L. A. et al. (2013). How visual edge features influence cuttlefish camouflage patterning. Vision Research, 83, 4047.Google Scholar
Chiao, C.-C., Wickiser, J. K., Allen, J. J., Genter, B. & Hanlon, R. T. (2011). Hyperspectral imaging of cuttlefish camouflage indicates good color match in the eyes of fish predators. Proceedings of the National Academy of Sciences, 108, 91489153.CrossRefGoogle ScholarPubMed
Chichery, M. P. & Chichery, R. (1992). Behavioral and neurohistological changes in aging Sepia. Brain Research, 574, 7784.Google Scholar
Chichery, R. & Chanelet, J. (1976). Motor and behavioural responses obtained by stimulation with chronic electrodes of the optic lobe of Sepia officinalis. Brain Research, 105, 525532.Google Scholar
Chiou, T. H., Mathger, L. M., Hanlon, R. T. & Cronin, T. W. (2007). Spectral and spatial properties of polarized light reflections from the arms of squid (Loligo pealeii) and cuttlefish (Sepia officinalis L.). Journal of Experimental Biology, 210, 36243635.Google Scholar
Chun, C. (1910). Die Cephalopoden. Wissenschaftliche Ergebnisse der Deutschen Tiefsee-Expedition auf dem Dampfer ‘Valdivia’ 1898–1899, 18, 1552.Google Scholar
Chun, C. (1914). Cephalopoda from the ‘Michael Sars’ North Atlantic Deep-Sea Expedition, 1910. Reports on Sars North Atlantic Deep Sea Expedition, 3, 128.Google Scholar
Cigliano, J. (1993). Dominance hierarchy and den use in Octopus bimaculoides. Animal Behaviour, 46, 677684.CrossRefGoogle Scholar
Cigliano, J. A. (1995). Assessment of the mating history of female pygmy octopuses and a possible sperm competition mechanism. Animal Behaviour, 49, 849851.Google Scholar
Clarke, A., Rodhouse, P. G. & Gore, D. J. (1994). Biochemical composition in relation to the energetics of growth and sexual maturation in the ommastrephid squid Illex argentinus. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 344, 201212.Google Scholar
Clarke, M. R. (1962). Significance of cephalopod beaks. Nature, London, 193, 560561.Google Scholar
Clarke, M. R. (1966). A review of the systematics and ecology of oceanic squids. Advances in Marine Biology, 4, 91300.CrossRefGoogle Scholar
Clarke, M. R. (1969). Cephalopoda collected on the SOND cruise. Journal of the Marine Biological Association of the United Kingdom, 49, 961976.Google Scholar
Clarke, M. R. (1977). Beaks, nets and numbers. Symposia of the Zoological Society of London, 38, 89126.Google Scholar
Clarke, M. R. (1980). Cephalopoda in the diet of sperm whales of the southern hemisphere and their bearing on sperm whale biology. Discovery Reports, 37, 1324.Google Scholar
Clarke, M. R. (1983). Cephalopod biomass – estimation from predation. Memoirs of the National Museum of Victoria, 44, 95107.Google Scholar
Clarke, M. R. (1985). Cephalopods in the diet of cetaceans and seals. Rapport Commission Internationale pour la Mer Mediterranee, 29, 211219.Google Scholar
Clarke, M. R. (1986). Handbook for the Identification of Cephalopod Beaks. London: Oxford University Press.Google Scholar
Clarke, M. R. (1988a). Evolution of buoyancy and locomotion in recent cephalopods, in The Mollusca, Vol. 12: Paleontology and Neontology of Cephalopods (ed. Clarke, M. R. & Trueman, E. R.), pp. 203213. San Diego: Academic Press.Google Scholar
Clarke, M. R. (1988b). Evolution of recent cephalopods – a brief review, in The Mollusca, Vol. 12: Paleontology and Neontology of Cephalopods (ed. Clarke, M. R. & Trueman, E. R.), pp. 331340. San Diego: Academic Press.Google Scholar
Clarke, M. R. (1996). Cephalopods as prey. 3. Cetaceans. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 351, 10531065.Google Scholar
Clarke, M. R. & Stevens, J. D. (1974). Cephalopods, blue sharks and migration. Journal of the Marine Biological Association of the United Kingdom, 54, 949957.Google Scholar
Clarke, M. R. & Trueman, E. R. (1988). The Mollusca, Vol. 12: Paleontology and Neontology of Cephalopods. San Diego, CA.: Academic Press.Google Scholar
Clarke, M. R., Denton, E. J. & Gilpin-Brown, J. B. (1979). Use of ammonium for buoyancy in squids. Journal of the Marine Biological Association of the United Kingdom, 59, 259276.Google Scholar
Cloney, R. A. & Brocco, S. L. (1983). Chromatophore organs, reflector cells, iridocytes and leucophores in cephalopods. American Zoologist, 23, 581592.Google Scholar
Cloney, R. A. & Florey, E. (1968). Ultrastructure of cephalopod chromatophore organs. Zeitschrift für Zellforschung, 89, 250280.Google Scholar
Clutton-Brock, T. H. (1988). Reproductive Success. Studies of Individual Variation in Contrasting Breeding Systems. Chicago and London: University of Chicago Press.Google Scholar
Clutton-Brock, T. H. (2009). Sexual selection in females. Animal Behaviour, 77, 311.Google Scholar
Cobb, C. S., Pope, S. K. & Williamson, R. (1995). Circadian rhythms to light–dark cycles in the lesser octopus, Eledone cirrhosa. Marine and Freshwater Behaviour and Physiology, 26, 4757.Google Scholar
Coelho, M. L. (1985). Review of the influence of oceanographic factors on cephalopod distribution and life cycles. NAFO Scientific Council Studies, 9, 4757.Google Scholar
Coelho, M. L., Quintela, J., Bettencourt, V., Olavo, G. & Villa, H. (1994). Population structure, maturation patterns and fecundity of the squid Loligo vulgaris from southern Portugal. Fisheries Research, 21, 87102.Google Scholar
Cohen, A. C. (1976). The systematics and distribution of Loligo (Cephalopoda, Myopsida) in the western North Atlantic, with descriptions of two new species. Malacologia, 15, 299367.Google Scholar
Cole, K. S. & Gilbert, D. L. (1970). Jet propulsion of squid. Biological Bulletin, 138, 245246.CrossRefGoogle Scholar
Cole, P. D. & Adamo, S. A. (2005). Cuttlefish (Sepia officinalis: Cephalopoda) hunting behavior and associative learning. Animal Cognition, 8, 2730.Google Scholar
Coleman, N. (1984). Molluscs from the diets of commercially exploited fish off the coast of Victoria, Australia. Journal of the Malacological Society of Australia, 6, 143154.CrossRefGoogle Scholar
Colgan, P. W. (1978). Quantitative Ethology. New York: John Wiley & Sons.Google Scholar
Collins, M. A., Burnell, G. M. & Rodhouse, P. (1995). Reproductive strategies of male and female Loligo forbesi (Cephalopoda: Loliginidae). Journal of the Marine Biological Association of the United Kingdom, 75, 621634.Google Scholar
Colmers, W. F., Hixon, R. F., Hanlon, R. T. et al. (1984). ‘Spinner’ cephalopods: Defects of statocyst suprastructures in an invertebrate analogue of the vestibular apparatus. Cell and Tissue Research, 236, 505515.Google Scholar
Condit, R. & Le Boeuf, B. J. (1984). Feeding habits and feeding grounds of the northern elephant seal. Journal of Mammalogy, 65, 281290.CrossRefGoogle Scholar
Cooper, K. M. & Hanlon, R. T. (1986). Correlation of iridescence with changes in iridophore platelet ultrastructure in the squid Lolliguncula brevis. Journal of Experimental Biology, 121, 451455.Google Scholar
Cooper, K. M., Hanlon, R. T. & Budelmann, B. U. (1990). Physiological color change in squid iridophores. II. Ultrastructural mechanisms in Lolliguncula brevis. Cell and Tissue Research, 259, 1524.Google Scholar
Corner, B. D. & Moore, H. T. (1980). Field observations on the reproductive behavior of Sepia latimanus. Micronesica, 16, 235260.Google Scholar
Cornwell, C. J., Messenger, J. B. & Hanlon, R. T. (1997). Chromatophores and body patterning in the squid Alloteuthis subulata. Journal of the Marine Biological Association of the United Kingdom, 77, 12431246.Google Scholar
Cortez, T., Castro, B. G. & Guerra, A. (1998). Drilling behaviour of Octopus mimus Gould. Journal of Experimental Marine Biology and Ecology, 224, 193203.Google Scholar
Cosgrove, J. (2003). An in situ observation of webover hunting by the Giant Pacific Octopus, Enteroctopus dofleini (Wulker, 1910). The Canadian Field-Naturalist, 117, 117118.Google Scholar
Cott, H. B. (1940). Adaptive Coloration in Animals. London: Methuen.Google Scholar
Cousteau, J. Y. & Diolé, P. (1973). Octopus and Squid: The Soft Intelligence. Garden City, NY: Doubleday.Google Scholar
Cowdry, E. V. (1911). The Colour Changes of Octopus vulgaris. University of Toronto Studies, Biological Series No. 10. Toronto: The University Library, The Librarian.Google Scholar
Cowen, R., Gertman, R. & Wiggett, G. (1973). Camouflage patterns in Nautilus and their implications for cephalopod paleobiology. Lethaia, 6, 201213.Google Scholar
Crancher, P., King, M. G., Bennett, A. & Montgomery, R. B. (1972). Conditioning of a free operant in Octopus cyanea Gray. Journal of the Experimental Analysis of Behavior, 17, 359362.Google Scholar
Crook, R. J. & Basil, J. A. (2008). A biphasic memory curve in the chambered nautilus, Nautilus pompilius L. (Cephalopoda: Nautiloidea). Journal of Experimental Biology, 211, 19921998.Google Scholar
Crook, R. J. & Basil, J. A. (2013). Flexible spatial orientation and navigational strategies in chambered Nautilus. Ethology, 119, 7785.Google Scholar
Crook, R. J. & Walters, E. T. (2011). Nociceptive behavior and physiology of molluscs: animal welfare implications. Ilar Journal, 52, 185195.Google Scholar
Crook, R. J., Dickson, K., Hanlon, R. T. & Walters, E. T. (2014). Nociceptive sensitization reduces predation risk. Current Biology, 24, 11211125.Google Scholar
Crook, R. J., Hanlon, R. T. & Basil, J. A. (2009). Memory of visual and topographical features suggests spatial learning in nautilus (Nautilus pompilius L.). Journal of Comparative Psychology, 123, 264274.Google Scholar
Crook, R. J., Hanlon, R. T. & Walters, E. T. (2013). Squid have nociceptors that display widespread long-term sensitization and spontaneous activity after bodily injury. Journal of Neuroscience, 33, 1002110026.Google Scholar
Crook, R. J., Lewis, T., Hanlon, R. T. & Walters, E. T. (2011). Peripheral injury induces long-term sensitization of defensive responses to visual and tactile stimuli in the squid Loligo pealeii, Lesueur 1821. Journal of Experimental Biology, 214, 31733185.Google Scholar
Crookes, W. J., Ding, L., Huang, Q. L. et al. (2004). Reflectins: the unusual proteins of squid reflective tissues. Science, 303, 235238.Google Scholar
Croxall, J. P. & Prince, P. A. (1996). Cephalopods as prey. 1. Seabirds. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 351, 10231043.Google Scholar
Cummins, S. F., Boal, J. G., Buresch, K. C. et al. (2011). Extreme aggression in male squid induced by a beta-MSP-like pheromone. Current Biology, 21, 322327.Google Scholar
Curio, E. (1976). The Ethology of Predation. Berlin: Springer-Verlag.Google Scholar
Dahlgren, H. (1916). The production of light by animals. Light production in cephalopods. Journal of the Franklin Institute, 525–556.Google Scholar
D’Aniello, A., DiCosmo, A., DiCristo, C. et al. (1996). Occurrence of sex steroid hormones and their binding proteins in Octopus vulgaris lam. Biochemical and Biophysical Research Communications, 227, 782788.Google Scholar
Darmaillacq, A. S. & Shashar, N. (2008). Lack of polarization optomotor response in the cuttlefish Sepia elongata (d’Orbigny, 1845). Physiology and Behavior, 94, 616620.Google Scholar
Darmaillacq, A. S., Chichery, R. & Dickel, L. (2006). Food imprinting, new evidence from the cuttlefish Sepia officinalis. Biology Letters, 2, 345347.Google Scholar
Darmaillacq, A. S., Chichery, R., Poirier, R. & Dickel, L. (2004a). Effect of early feeding experience on subsequent prey preference by cuttlefish, Sepia officinalis. Developmental Psychobiology, 45, 239244.Google Scholar
Darmaillacq, A. S., Chichery, R., Shashar, N. & Dickel, L. (2006). Early familiarization overrides innate prey preference in newly hatched Sepia officinalis cuttlefish. Animal Behaviour, 71, 511514.Google Scholar
Darmaillacq, A. S., Dickel, L., Chichery, M. P., Agin, V. & Chichery, R. (2004b). Rapid taste aversion learning in adult cuttlefish, Sepia officinalis. Animal Behaviour, 68, 12911298.Google Scholar
Darmaillacq, A. S., Dickel, L. & Mather, J. (2014). Cephalopod Cognition. Cambridge: Cambridge University Press.Google Scholar
Darmaillacq, A. S., Lesimple, C. & Dickel, L. (2008). Embryonic visual learning in the cuttlefish, Sepia officinalis. Animal Behaviour, 76, 131134.Google Scholar
Darwin, C. (1871). The Descent of Man and Selection in Relation to Sex. London: Murray.Google Scholar
Davey, G. (1989). Ecological Learning Theory. London: Routledge.Google Scholar
Davies, N. B. (1991). Mating systems, in Behavioural Ecology. An Evolutionary Approach (ed. Krebs, J. R. & Davies, N. B.), pp. 263294. Oxford: Blackwell Scientific Publications.Google Scholar
Davis, R. W., Jaquet, N., Gendron, D. et al. (2007). Diving behavior of sperm whales in relation to behavior of a major prey species, the jumbo squid, in the Gulf of California, Mexico. Marine Ecology Progress Series, 333, 291302.Google Scholar
Daw, N. W. & Pearlman, A. L. (1974). Pigment migration and adaptation in eye of squid, Loligo pealei. Journal of General Physiology, 63, 2236.Google Scholar
Dawkins, M. (1971). Perceptual changes in chicks: another look at the ‘search image’ concept. Animal Behaviour, 19, 566574.Google Scholar
Dawkins, M. S. (2007). Observing Animal Behaviour: Design and Analysis of Quantitive Controls. Oxford: Oxford University Press.Google Scholar
Dawkins, M. S. & Guilford, T. (1991). The corruption of honest signalling. Animal Behaviour, 41, 865873.Google Scholar
Dawkins, R. & Krebs, J. R. (1978). Animal signals: information or manipulation?, in Behavioural Ecology: An Evolutionary Approach (ed. Krebs, J. R. & Davies, N. B.), pp. 282309. Sutherland, MA: Sinauer Associates, Inc.Google Scholar
de Beer, C. L. & Potts, W. M. (2013). Behavioural observations of the common octopus Octopus vulgaris in Baia dos Tigres, southern Angola. African Journal of Marine Science, 35, 579583.Google Scholar
DeLong, R. L., Kooyman, G. L., Gilmartin, W. G. & Loughlin, T. R. (1984). Hawaiian monk seal diving behavior. Proceedings of the Third International Theriological Congress, Helsinki; Acta Zoologica Fennica, 172, 129131.Google Scholar
DeMartini, D. G., Krogstad, D. V. & Morse, D. E. (2013). Membrane invaginations facilitate reversible water flux driving tunable iridescence in a dynamic biophotonic system. Proceedings of the National Academy of Sciences, 110, 25522556.CrossRefGoogle Scholar
Denton, E. J. (1970). On the organization of reflecting surfaces in some marine animals. Philosophical Transactions of the Royal Society of London B, 258, 285313.Google Scholar
Denton, E. J. (1974). On buoyancy and the lives of modern and fossil cephalopods. Proceedings of the Royal Society of London B, 185, 273299.Google Scholar
Denton, E. J. (1990). Light and vision at depths greater than 200 metres, in Light and Life in the Sea (ed. Herring, P. J., Campbell, A. K., Whitfield, M. & Maddock, L.), pp. 127148. Cambridge: Cambridge University Press.Google Scholar
Denton, E. J. & Gilpin-Brown, J. B. (1961). Effect of light on buoyancy of cuttlefish. Journal of the Marine Biological Association of the United Kingdom, 41, 343350.Google Scholar
Denton, E. J. & Gilpin-Brown, J. B. (1966). On the buoyancy of the pearly nautilus. Journal of the Marine Biological Association of the United Kingdom, 46, 723759.Google Scholar
Denton, E. J. & Gilpin-Brown, J. B. (1973). Floatation mechanisms in modern and fossil cephalopods. Advances in Marine Biology, 11, 197268.Google Scholar
Denton, E. J. & Land, M. F. (1971). Mechanism of reflexion in silvery layers of fish and cephalopods. Proceedings of the Royal Society of London B: Biological Sciences, 178, 4361.Google Scholar
Denton, E. J. & Locket, N. A. (1989). Possible wavelength discrimination by multibank retinae in deep-sea fish. Journal of the Marine Biological Association of the United Kingdom, 69, 409435.Google Scholar
Denton, E. J. & Warren, F. J. (1968). Eyes of the Histioteuthidae. Nature, 219, 400401.Google Scholar
Denton, E. J., Gilpin-Brown, J. B. & Howarth, J. V. (1967). On the buoyancy of Spirula spirula. Journal of the Marine Biological Association of the United Kingdom, 47, 181191.Google Scholar
Deravi, L. F., Magyar, A. P., Sheehy, S. P. et al. (2014). The structure–function relationships of a natural nanoscale photonic device in cuttlefish chromatophores. Journal of the Royal Society Interface, 11, 9.Google Scholar
Derby, C. D. (2007). Escape by inking and secreting: marine molluscs avoid predators through a rich array of chemicals and mechanisms. Biological Bulletin, 213, 274289.Google Scholar
Derby, C. D. (2014). Cephalopod ink: production, chemistry, functions and applications. Marine Drugs, 12, 27002730.Google Scholar
Derby, C. D., Kicklighter, C. E., Johnson, P. M. & Zhang, X. (2007). Chemical composition of inks of diverse marine molluscs suggests convergent chemical defenses. Journal of Chemical Ecology, 33, 11051113.Google Scholar
Derby, C. D., Tottempudi, M., Love-Chezem, T. & Wolfe, L. S. (2013). Ink from longfin inshore squid, Doryteuthis pealeii, as a chemical and visual defense against two predatory fishes, summer flounder, Paralichthys dentatus, and sea catfish, Ariopsis felis. Biological Bulletin, 225, 152160.CrossRefGoogle ScholarPubMed
Dews, P. M. (1959). Some observations on an operant in the octopus. Journal of the Experimental Analysis of Behavior, 2, 5763.Google Scholar
Diamant, A. & Shpigel, M. (1985). Interspecific feeding associations of groupers (Teleostei: Serranidae) with octopuses and moray eels in the Gulf of Eilat (Aqaba). Environmental Biology of Fishes, 13, 153159.Google Scholar
Dickel, L., Chichery, M. P. & Chichery, R. (1997). Postembryonic maturation of the vertical lobe complex and early development of predatory behavior in the cuttlefish (Sepia officinalis). Neurobiology of Learning and Memory, 67, 150160.Google Scholar
Dickel, L., Chichery, M. P. & Chichery, R. (1998). Time differences in the emergence of short- and long-term memory during post-embryonic development in the cuttlefish, Sepia. Behavioural Processes, 44, 8186.Google Scholar
Dickel, L., Chichery, M. P. & Chichery, R. (2001). Increase of learning abilities and maturation of the vertical lobe complex during postembryonic development in the cuttlefish, Sepia. Developmental Psychobiology, 39, 9298.Google Scholar
Di Cosmo, A. & Di Cristo, C. (1998). Neuropeptidergic control of the optic gland of Octopus vulgaris: FMRF-amide and GnRH immunoreactivity. Journal of Comparative Neurology, 398, 112.Google Scholar
Di Cosmo, A., Di Cristo, C. & Paolucci, M. (2001). Sex steroid hormone fluctuations and morphological changes of the reproductive system of the female of Octopus vulgaris throughout the annual cycle. Journal of Experimental Zoology, 289, 3347.Google Scholar
Di Cristo, C. (2013). Nervous control of reproduction in Octopus vulgaris: a new model. Invertebrate Neuroscience, 13, 2734.CrossRefGoogle ScholarPubMed
Dijkgraaf, S. (1961). The statocyst of Octopus vulgaris as a rotation receptor. Pubblicazioni della Stazione Zoologica di Napoli, 32, 6487.Google Scholar
Dijkgraaf, S. (1963). Versuche über Schallwahrnehmung bei Tintenfischen. Naturwissenschaften, 50, 50.Google Scholar
Dilly, P. N. (1963). Delayed responses in Octopus. Journal of Experimental Biology, 40, 393401.CrossRefGoogle Scholar
Dilly, P. N. (1972). Taonius megalops, a squid that rolls up into a ball. Nature, 237, 403404.Google Scholar
Dilly, P. N. & Herring, P. J. (1978). The light organ and ink sac of Heteroteuthis dispar (Mollusca: Cephalopoda). Journal of Zoology (London), 186, 4759.Google Scholar
Dilly, P. N. & Herring, P. J. (1981). Ultrastructural features of the light organs of Histioteuthis macrohista (Mollusca, Cephalopoda). Journal of Zoology, 195, 255266.Google Scholar
Dilly, P. N., Nixon, M. & Packard, A. (1964). Forces exerted by Octopus vulgaris. Pubblicazioni della Stazione Zoologica di Napoli, 34, 8697.Google Scholar
DiMarco, F. P. & Hanlon, R. T. (1997). Agonistic behavior in the squid Loligo plei (Loliginidae, Teuthoidea): fighting tactics and the effects of size and resource value. Ethology, 103, 89108.Google Scholar
Dimitrova, M., Stobbe, N., Schaefer, H. M. & Merilaita, S. (2009). Concealed by conspicuousness: distractive prey markings and backgrounds. Proceedings of the Royal Society B – Biological Sciences, 276, 19051910.Google Scholar
Domingues, P., Sykes, A., Sommerfield, A. & Andrade, J. P. (2003). Effects of feeding live or frozen prey on growth, survival and the life cycle of the cuttlefish, Sepia officinalis (Linnaeus, 1758). Aquaculture International, 11, 397410.Google Scholar
Donovan, D. T. (1964). Cephalopod phylogeny and classification. Biological Reviews, 39, 259287.Google Scholar
d’Orbigny, A. (1841). Histoire naturelle générale et particulière des Céphalopodes. Paris.Google Scholar
Dorsey, E. M. (1976). Natural history and social behavior of Octopus rubescens Berry. Seattle: Unpublished M.S. Thesis, University of Washington.Google Scholar
Doubleday, Z. A., Semmens, J. M., Smolenski, A. J. & Shaw, P. W. (2009). Microsatellite DNA markers and morphometrics reveal a complex population structure in a merobenthic octopus species (Octopus maorum) in south-east Australia and New Zealand. Marine Biology, 156, 11831192.CrossRefGoogle Scholar
Douglas, R. H., Williamson, R. & Wagner, H. J. (2005). The pupillary response of cephalopods. Journal of Experimental Biology, 208, 261265.Google Scholar
Drazen, J. C., Goffredi, S. K., Schlining, B. & Stakes, D. S. (2003). Aggregations of egg-brooding deep-sea fish and cephalopods on the Gorda Escarpment: a reproductive hot spot. Biological Bulletin, 205, 17.Google Scholar
Drew, G. A. (1911). Sexual activities of the squid, Loligo pealii (Les.). Journal of Morphology, 22, 327359.Google Scholar
Driver, P. M. & Humphries, D. A. (1988). Protean Behaviour: The Biology of Unpredictability. New York: Oxford University Press.Google Scholar
Drummond, H. (1981). The nature and description of behavior patterns, in Perspectives in Ethology, Vol. 4: Advantages of Diversity (ed. Bateson, P. P. G. & Klopfer, P. H.), pp. 133. New York and London: Plenum Press.Google Scholar
Dubas, F. & Boyle, P. R. (1985). Chromatophore motor units in Eledone cirrhosa (Cephalopoda: Octopoda). Journal of Experimental Biology, 117, 415432.Google Scholar
Dubas, F., Hanlon, R. T., Ferguson, G. P. & Pinsker, H. M. (1986). Localization and stimulation of chromatophore motoneurones in the brain of the squid, Lolliguncula brevis. Journal of Experimental Biology, 121, 125.Google Scholar
Dubas, F., Leonard, R. B. & Hanlon, R. T. (1986). Chromatophore motoneurons in the brain of the squid, Lolliguncula brevis: an HRP study. Brain Research, 374, 2129.Google Scholar
Dunstan, A., Alanis, O. & Marshall, J. (2010). Nautilus pompilius fishing and population decline in the Philippines: a comparison with an unexploited Australian Nautilus population. Fisheries Research, 106, 239247.Google Scholar
Dunstan, A., Bradshaw, C. J. A. & Marshall, J. (2011). Nautilus at risk – estimating population size and demography of Nautilus pompilius. PLoS One, 6, e16716.Google Scholar
Dunstan, A. J., Ward, P. D. & Marshall, N. J. (2011a). Nautilus pompilius life history and demographics at the Osprey Reef Seamount, Coral Sea, Australia. PLoS One, 6, e16312.Google Scholar
Dunstan, A. J., Ward, P. D. & Marshall, N. J. (2011b). Vertical distribution and migration patterns of Nautilus pompilius. PLoS One, 6, e16311.Google Scholar
Durward, R. D., Vessey, E., O’Dor, R. K. & Amaratunga, T. (1980). Reproduction in the squid, Illex illecebrosus: first observations in captivity and implications for the life cycle. International Commission for the Northwest Atlantic Fisheries Selected Papers, 6, 713.Google Scholar
Duval, P., Chichery, M.-P. & Chichery, R. (1984). Prey capture by the cuttlefish (Sepia officinalis L.): an experimental study of two strategies. Behavioural Processes, 9, 1321.Google Scholar
Eberhard, W. G. (1996). Female Control: Sexual Selection by Cryptic Female Choice. New Jersey: Princeton University Press.Google Scholar
Eberhard, W. G. (2000). Criteria for demonstrating postcopulatory female choice. Evolution, 54, 10471050.Google Scholar
Eberhard, W. G. (2009). Postcopulatory sexual selection: Darwin’s omission and its consequences. Proceedings of the National Academy of Sciences of the United States of America, 106, 1002510032.Google Scholar
Eberhard, W. G. & Cordero, C. (2003). Sexual conflict and female choice. Trends in Ecology & Evolution, 18, 438439.Google Scholar
Ebisawa, S., Tsuchiya, K. & Segawa, S. (2011). Feeding behavior and oxygen consumption of Octopus ocellatus preying on the short-neck clam Ruditapes philippinarum. Journal of Experimental Marine Biology and Ecology, 403, 18.Google Scholar
Edelman, D. B. & Seth, A. K. (2009). Animal consciousness: a synthetic approach. Trends in Neurosciences, 32, 476484.Google Scholar
Edmunds, M. (1974). Defence in Animals. A Survey of Anti-Predator Defences. New York: Longman Group, Ltd.Google Scholar
Edmunds, M. (2000). Why are there good and poor mimics? Biological Journal of the Linnean Society, 459–466.Google Scholar
Edut, S. & Eilam, D. (2004). Protean behavior under barn-owl attack: voles alternate between freezing and fleeing and spiny mice flee in alternating patterns. Behavioural Brain Research, 155, 207216.Google Scholar
Ehrhardt, N. M. (1991). Potential impact of a seasonal migratory jumbo squid (Dosidicus gigas) stock on a Gulf of California sardine (Sardinops sagax caerulea) population. Bulletin of Marine Science, 49, 325332.Google Scholar
Eibl-Eibesfeldt, I. (1975). Ethology. The Biology of Behavior. New York: Holt, Rinehart and Winston, Inc.Google Scholar
Eibl-Eibesfeldt, I. & Scheer, G. (1962). Das brutpflegeverhalten eines weiblichen Octopus aegina Gray. Zeitschrift für Tierpsychologie, 19, 257261.Google Scholar
Ellis, D. V. (1985). Animal Behavior and its Applications. Chelsea, MI: Lewis Publishers, Inc.Google Scholar
Ellis, R. (1999). The Search for the Giant Squid. Robert Hale, Ltd.Google Scholar
Elwood, R. W. & Arnott, G. (2012). Understanding how animals fight with Lloyd Morgan’s canon. Animal Behaviour, 84, 10951102.Google Scholar
Emery, A. M., Wilson, I. J., Craig, S., Boyle, P. R. & Noble, L. R. (2001). Assignment of paternity groups without access to parental genotypes: multiple mating and developmental plasticity in squid. Molecular Ecology, 10, 12651278.Google Scholar
Emery, D. G. (1975). The histology and fine structure of the olfactory organ in the squid Lolliguncula brevis Blainville. Tissue and Cell, 7, 357367.Google Scholar
Emery, D. G. (1976). Observations on the olfactory organ of adult and juvenile Octopus joubini. Tissue and Cell, 8, 3346.Google Scholar
Emlen, S. T. & Oring, L. W. (1977). Ecology, sexual selection and the evolution of mating systems. Science, 197, 215223.Google Scholar
Enault, J., Zatylny-Gaudin, C., Bernay, B. et al. (2012). A complex set of sex pheromones identified in the cuttlefish Sepia officinalis. PLoS ONE, 7, e46531.Google Scholar
Endler, J. A. (1978). A predator’s view of animal color patterns. Evolutionary Biology, 11, 319364.Google Scholar
Endler, J. A. (1981). An overview of the relationships between mimicry and crypsis. Biological Journal of the Linnean Society, 16, 2531.Google Scholar
Endler, J. A. (1984). Progressive background matching in moths, and a quantitative measure of crypsis. Biological Journal of the Linnean Society, 22, 187231.Google Scholar
Endler, J. A. (1991). Interactions between predators and prey, in Behavioural Ecology. An Evolutionary Approach (ed. Krebs, J. R. & Davies, N. B.), pp. 169196. Oxford: Blackwell Scientific Publications.Google Scholar
Endler, J. A. (1993). Some general comments on the evolution and design of animal communication systems. Philos Trans R Soc Lond B Biol Sci, 340, 215225.Google Scholar
Endler, J. A. (2006). Disruptive and cryptic coloration. Proceedings of the Royal Society B – Biological Sciences, 273, 24252426.Google Scholar
Endler, J. A. & Mielke, P. W. (2005). Comparing entire colour patterns as birds see them. Biological Journal of the Linnean Society, 86, 405431.Google Scholar
Engeser, T. S. & Clarke, M. R. (1988). Cephalopod hooks, both recent and fossil, in The Mollusca, Vol. 12: Paleontology and Neontology of Cephalopods (ed. Clarke, M. R. & Trueman, E. R.), pp. 133151. San Diego: Academic Press.Google Scholar
Englund, G. & Olsson, T. I. (1990). Fighting and assessment in the net-spinning caddis larva Arctopsyche ladogensis: a test of the sequential assessment game. Animal Behaviour, 39, 5562.Google Scholar
Enquist, M. & Leimar, O. (1983). Evolution of fighting behavior: decision rules and assessment of relative strength. Journal of Theoretical Biology, 102, 387410.Google Scholar
Enquist, M. & Leimar, O. (1987). Evolution of fighting behaviour: the effect of variation in resource value. Journal of Theoretical Biology, 127, 187205.Google Scholar
Enquist, M., Ljungberg, T. & Zandor, A. (1987). Visual assessment of fighting ability in the Cichlid fish Nannacara anomala. Animal Behaviour, 35, 12621263.Google Scholar
Escanez, A., Riera, R., Gonzalez, A. F. & Guerra, A. (2012). On the occurrence of egg masses of the diamond-shaped squid Thysanoteuthis rhombus Troschel, 1857 in the subtropical eastern Atlantic (Canary Islands). A potential commercial species? ZooKeys, 222, 6976.Google Scholar
Eyman, M., Crispino, M., Kaplan, B. B. & Giuditta, A. (2003). Squid photoreceptor terminals synthesize calexcitin, a learning related protein. Neuroscience Letters, 347, 2124.Google Scholar
Fawcett, T. W. & Mowles, S. L. (2013). Assessments of fighting ability need not be cognitively complex. Animal Behaviour, 86, E1E7.Google Scholar
Ferguson, G. P. & Messenger, J. B. (1991). A countershading reflex in cephalopods. Proceedings of the Royal Society of London B: Biological Sciences, 243, 6367.Google Scholar
Ferguson, G. P., Martini, F. M. & Pinsker, H. M. (1988). Chromatophore motor fields in the squid, Lolliguncula brevis. Journal of Experimental Biology, 134, 281195.Google Scholar
Ferguson, G. P., Messenger, J. B. & Budelmann, B. U. (1994). Gravity and light influence the countershading reflexes of the cuttlefish Sepia officinalis. Journal of Experimental Biology, 191, 247256.Google Scholar
Field, J. C., Elliger, C., Baltz, K. et al. (2013). Foraging ecology and movement patterns of jumbo squid (Dosidicus gigas) in the California Current System. Deep-Sea Research Part II – Topical Studies in Oceanography, 95, 3751.Google Scholar
Fields, W. G. (1965). The structure, development, food relations, reproduction, and life history of the squid Loligo opalescens Berry. Fishery Bulletin, 131, 1108.Google Scholar
Finn, J. K. & Norman, M. D. (2010). The argonaut shell: gas-mediated buoyancy control in a pelagic octopus. Proceedings of the Royal Society B – Biological Sciences, 277, 29672971.Google Scholar
Finn, J. K., Tregenza, T. & Norman, M. D. (2009a). Defensive tool use in a coconut-carrying octopus. Current Biology, 19, R1069R1070.Google Scholar
Finn, J., Tregenza, T. & Norman, M. (2009b). Preparing the perfect cuttlefish meal: complex prey handling by dolphins. PLoS ONE, 4, e4217.Google Scholar
Fiorito, G. & Chichery, R. (1995). Lesions of the vertical lobe impair visual discrimination learning by observation in Octopus vulgaris. Neuroscience Letters, 192, 117120.Google Scholar
Fiorito, G. & Gherardi, F. (1999). Prey-handling behaviour of Octopus vulgaris (Mollusca, Cephalopoda) on bivalve preys. Behavioural Processes, 46, 7588.Google Scholar
Fiorito, G. & Scotto, P. (1992). Observational learning in Octopus vulgaris. Science, 256, 545547.Google Scholar
Fiorito, G., Agnisola, C., d’Addio, M., Valanzano, A. & Calamandrei, G. (1998). Scopolamine impairs memory recall in Octopus vulgaris. Neuroscience Letters, 253, 8790.Google Scholar
Fiorito, G., von Planta, C. & Scotto, P. (1990). Problem solving ability of Octopus vulgaris Lamarck (Mollusca, Cephalopoda). Behavioral and Neural Biology, 53, 217230.Google Scholar
Fleisher, K. J. & Case, J. F. (1995). Cephalopod predation facilitated by dinoflagellate luminescence. Biological Bulletin, 189, 263271.Google Scholar
Fleming, I. (1958). Dr No. London: Cape.Google Scholar
Flores, E. E. C. (1983). Visual discrimination testing in the squid Todarodes pacificus: experimental evidence for lack of color vision. Memoirs of the National Museum of Victoria, 44, 213227.Google Scholar
Flores, E. E. C., Igarashi, S. & Mikami, T. (1978). Studies on squid behavior in relation to fishing – III. On the optomotor response of squid Todarodes pacificus Steenstrup, to various colors. Bulletin of the Faculty of Fisheries, Hokkaido University, 29, 131140.Google Scholar
Florey, E. (1966). Nervous control and spontaneous activity of the chromatophores of a cephalopod, Loligo opalescens. Comparative Biochemistry and Physiology, 18, 305324.Google Scholar
Florey, E. (1969). Ultrastructure and function of cephalopod chromatophores. American Zoologist, 9, 429442.Google Scholar
Florey, E. (1985). The Zoological Station at Naples and the neuron: personalities and encounters in a unique institution. Biological Bulletin, 168, 137152.Google Scholar
Florey, E. & Kriebel, M. E. (1969). Electrical and mechanical responses of chromatophore muscle fibers of squid, Loligo opalescens, to nerve stimulation and drugs. Zeitschrift für Vergleichende Physiologie, 65, 98130.Google Scholar
Florey, E., Dubas, F. & Hanlon, R. T. (1985). Evidence for L-glutamate as a transmitter substance of motoneurons innervating squid chromatophore muscles. Comparative Biochemistry and Physiology C, 82, 259268.Google Scholar
Foote, K. G., Hanlon, R. T., Iampietro, P. J. & Kvitek, R. G. (2006). Acoustic detection and quantification of benthic egg beds of the squid Loligo opalescens in Monterey Bay, California. Journal of the Acoustic Society of America, 119, 844856.Google Scholar
Ford, E. B. (1975). Ecological Genetics. London: Chapman & Hall.Google Scholar
Forsythe, J. W. & Hanlon, R. T. (1985). Aspects of egg development, post-hatching behavior, growth and reproductive biology of Octopus burryi Voss, 1950 (Mollusca: Cephalopoda). Vie Milieu, 35, 273282.Google Scholar
Forsythe, J. W. & Hanlon, R. T. (1988). Behavior, body patterning and reproductive biology of Octopus bimaculoides from California. Malacologia, 29, 4155.Google Scholar
Forsythe, J. W. & Hanlon, R. T. (1997). Foraging and associated behavior by Octopus cyanea Gray, 1849 on a coral atoll, French Polynesia. Journal of Experimental Marine Biology and Ecology, 209, 1531.Google Scholar
Forsythe, J. W. & Van Heukelem, W. F. (1987). Growth, in Cephalopod Life Cycles, Vol. 2: Comparative Reviews (ed. Boyle, P. R.), pp. 135156. London: Academic Press.Google Scholar
Forsythe, J. W., DeRusha, R. H. & Hanlon, R. T. (1994). Growth, reproduction and life-span of Sepia officinalis (Cephalopoda, Mollusca) cultured through seven consecutive generations. Journal of Zoology, 233, 175192.Google Scholar
Forsythe, J. W., Kangas, N. & Hanlon, R. T. (2004). Does the California market squid, Loligo opalescens, spawn naturally during the day or at night? A note on the successful use of ROV’s to obtain basic fisheries biology data. Fisheries Bulletin, 102, 389392.Google Scholar
Fotheringham, N. (1974). Tropic complexity in a littoral boulderfield. Limnology and Oceanography, 19, 8491.Google Scholar
Fox, D. L. (1938). An illustrated note on the mating and egg-brooding habits of the two-spotted octopus. Transactions of the San Diego Society of Natural History, 9, 3134.Google Scholar
Foyle, T. P. & O’Dor, R. K. (1987). Predatory strategies of squid (Illex illecebrosus) attacking small and large fish. Marine Behaviour and Physiology, 13, 155168.Google Scholar
Francisco Ruiz, J., Sepulveda, R. D. & Ibanez, C. M. (2012). Behaviour of Robsonella fontaniana in response to a potential predator. Latin American Journal of Aquatic Research, 40, 253258.Google Scholar
Frank, M. G., Waldrop, R. H., Dumoulin, M., Aton, S. & Boal, J. G. (2012). A preliminary analysis of sleep-like states in the cuttlefish Sepia officinalis. PLoS ONE, 7, e38125.Google Scholar
Franklin, A. M., Squires, Z. E. & Stuart-Fox, D. (2012). The energetic cost of mating in a promiscuous cephalopod. Biology Letters, 8, 754756.Google Scholar
Froesch, D. (1973). Projection of chromatophore nerves on the body surface of Octopus vulgaris. Marine Biology, 19, 203242.Google Scholar
Froesch, D. & Marthy, H.-J. (1975). The structure and function of the oviducal gland in octopods (Cephalopoda). Proceedings of the Royal Society of London B, 188, 95101.Google Scholar
Froesch, D. & Messenger, J. B. (1978). On leucophores and the chromatic unit of Octopus vulgaris. Journal of Zoology (London), 186, 163173.Google Scholar
Fukuda, Y. (1980). Observations by SEM, in Nautilus macromphalus in Captivity (ed. Hamada, T., Obata, I. & Okutani, T.), pp. 2333. Tokyo: Tokai University Press.Google Scholar
Gabe, S. H. (1975). Reproduction in the giant octopus of the North Pacific, Octopus dofleini martini. Veliger, 18, 146150.Google Scholar
Gadgil, M. & Bossert, W. H. (1970). Life historical consequences of natural selection. The American Naturalist, 104, 124.Google Scholar
Garcia-Gonzalez, F. & Simmons, L. W. (2007). Shorter sperm confer higher competitive fertilization success. Evolution, 61, 816824.Google Scholar
Gaston, M.R. & Tublitz, N.J. (2004). Peripheral innervation patterns and central distribution of fin chromatophore motoneurons in the cuttlefish Sepia officinalis. Journal of Experimental Biology 207, 30893098.Google Scholar
Gauldie, R. W., West, I. F. & Forch, E. C. (1994). Statocyst, statolith, and age estimation of the giant-squid, Architeuthis kirki. Veliger, 37, 93109.Google Scholar
Gennaro, J. F. J., Lorincz, A. E. & Brewster, H. B. (1965). The anterior gland of the octopus (Octopus vulgaris) and its mucous secretion. Annals of the New York Academy of Sciences, 118, 10211025.CrossRefGoogle ScholarPubMed
Gerlach, G., Buresch, K. C. & Hanlon, R. T. (2012). Population structure of the squid Doryteuthis (Loligo) pealeii on the eastern coast of the USA: comment on Shaw et al. (2010). Marine Ecology Progress Series, 450, 281283.Google Scholar
Ghiretti, F. (1959). Cephalotoxin: the crab-paralysing agent of the posterior salivary glands of cephalopods. Nature, 183, 11921193.Google Scholar
Ghiretti, F. (1960). Toxicity of octopus saliva against crustacea. Annals of the New York Academy of Sciences, 90, 726741.Google Scholar
Ghiretti, F. & Cariello, L. (1977). Gli Animali Marini Velenosi e le loro Tossine. Padova: Piccin.Google Scholar
Ghoshal, A., DeMartini, D. G., Eck, E. & Morse, D. E. (2013). Optical parameters of the tunable Bragg reflectors in squid. Journal of the Royal Society Interface, 10, 20130386.Google Scholar
Gibson, R. M. (1983). Visual abilities and foraging behaviour of predatory fish. Trends in Neuroscience, 6, 197199.Google Scholar
Gilly, W. F. & Lucero, M. T. (1992). Behavioural responses to chemical stimulation of the olfactory organ in the squid, Loligo opalescens. Journal of Experimental Biology, 162, 209229.Google Scholar
Gilly, W. F., Elliger, C. A., Salinas, C. A. et al. (2006). Spawning by jumbo squid Dosidicus gigas in San Pedro Martir basin, Gulf of California, Mexico. Marine Ecology Progress Series, 313, 125133.Google Scholar
Gilly, W. F., Hopkins, B. & Mackie, G. O. (1991). Development of giant motor axons and neural control of escape responses in squid embryos and hatchlings. Biological Bulletin, 180, 209220.Google Scholar
Gilly, W. F., Markaida, U., Baxter, C. H. et al. (2006). Vertical and horizontal migrations by the jumbo squid Dosidicus gigas revealed by electronic tagging. Marine Ecology – Progress Series, 324, 117.Google Scholar
Gilly, W. F., Zeidberg, L. D., Booth, J. A. T. et al. (2012). Locomotion and behavior of Humboldt squid, Dosidicus gigas, in relation to natural hypoxia in the Gulf of California, Mexico. The Journal of Experimental Biology, 215, 31753190.Google Scholar
Glanzman, D. L. (2008). Octopus conditioning: a multi-armed approach to the LTP-learning question. Current Biology, 18, R527R530.Google Scholar
Gleadall, I. G. & Shashar, N. (2004). The octopus’s garden: the visual world of cephalopods, in Complex Worlds from Simpler Nervous Systems (ed. Prete, F. R.), pp. 269308. Cambridge, MA: MIT Press.Google Scholar
Gleadall, I. G., Guerrero-Kommritz, J., Hochberg, F. G. & Laptikhovsky, V. V. (2010). The inkless octopuses (Cephalopoda: Octopodidae) of the southwest Atlantic. Zoological Science, 27, 528553.Google Scholar
Gleadall, I. G., Ohtsu, K., Gleadall, E. & Tsukahara, Y. (1993). Screening-pigment migration in the octopus retina includes control by dopaminergic efferents. Journal of Experimental Biology, 185, 116.Google Scholar
Godfrey-Smith, P. & Lawrence, M. (2012). Long-term high-density occupation of a site by Octopus tetricus and possible site modification due to foraging behavior. Marine and Freshwater Behaviour and Physiology, 45, 261268.Google Scholar
Gonzalez-Bellido, P. T., Wardill, T. J., Buresch, K. C., Ulmer, K. M. & Hanlon, R. T. (2014). Expression of squid iridescence depends on environmental luminance and peripheral ganglion control. Journal of Experimental Biology, 217, 850858.Google Scholar
Goss, C., Middleton, D. & Rodhouse, P. (2001). Investigations of squid stocks using acoustic survey methods. Fisheries Research, 54, 111121.Google Scholar
Grable, M. M., Shashar, N., Gilles, N. L., Chiao, C. C. & Hanlon, R. T. (2002). Cuttlefish body patterns as a behavioral assay to determine polarization perception. Biological Bulletin, 203, 232234.Google Scholar
Grasso, F. W. (2008). Octopus sucker-arm coordination in grasping and manipulation. American Malacological Bulletin, 24, 1323.Google Scholar
Grasso, F. W. & Basil, J. A. (2009). The evolution of flexible behavioral repertoires in cephalopod molluscs. Brain, Behavior and Evolution, 74, 231245.Google Scholar
Gratwicke, B. & Speight, M. R. (2005). The relationship between fish species richness, abundance and habitat complexity in a range of shallow tropical marine habitats. Journal of Fish Biology, 66, 650667.Google Scholar
Graziadei, P. (1964a). Electron microscopy of some primary receptors in the sucker of Octopus vulgaris. Zeitschrift für Zellforschung, 64, 510522.Google Scholar
Graziadei, P. (1964b). Receptors in the sucker of the cuttlefish. Nature, 203, 384386.Google Scholar
Graziadei, P. (1965). Sensory receptor cells and related neurons in cephalopods. Cold Spring Harbor Symposia on Quantitative Biology, 30, 4557.Google Scholar
Graziadei, P. (1971). The nervous system of the arms, in The Anatomy of the Nervous System of Octopus vulgaris (ed. Young, J. Z.), pp. 4561. Oxford: Clarendon.Google Scholar
Graziadei, P. P. C. & Gagne, H. T. (1976a). An unusual receptor in octopus. Tissue and Cell, 8, 229240.Google Scholar
Graziadei, P. P. C. & Gagne, H. T. (1976b). Sensory innervation in the rim of the octopus sucker. Journal of Morphology, 150, 639679.Google Scholar
Grimpe, G. (1926). Biologische Beobachtungen an Sepia officinalis. Deutsche Zoologische Gesellschaft Verhandlungen, 31, 148153.Google Scholar
Grisley, M. S. (1993). Separation and partial characterization of salivary enzymes expressed during prey handling in the octopus Eledone cirrhosa. Comparative Biochemistry and Physiology, 105B, 183192.Google Scholar
Grisley, M. S. & Boyle, P. R. (1985). A new application of serological techniques to gut content analysis. Journal of Experimental Marine Biology and Ecology, 90, 19.Google Scholar
Grisley, M. S. & Boyle, P. R. (1987). Bioassay and proteolytic activity of digestive enzymes from octopus saliva. Comparative Biochemistry and Physiology, 88B, 11171124.Google Scholar
Grisley, M. S. & Boyle, P. R. (1988). Recognition of food in Octopus digestive tract. Journal of Experimental Marine Biology and Ecology, 118, 732.Google Scholar
Grisley, M. S. & Boyle, P. R. (1990). Chitinase: a novel enzyme from octopus saliva. Comparative Biochemistry and Physiology, 95B, 311316.Google Scholar
Grisley, M. S., Boyle, P. R. & Key, L. N. (1996). Eye puncture as a route of entry for saliva during predation on crabs by the octopus Eledone cirrhosa (Lamarck). Journal of Experimental Marine Biology and Ecology, 202, 225237.Google Scholar
Griswold, C. A. & Prezioso, J. (1981). In situ observations on reproductive behavior of the long-finned squid, Loligo pealei. Fishery Bulletin, 78, 945947.Google Scholar
Groeger, G., Cotton, P. A. & Williamson, R. (2005). Ontogenetic changes in the visual acuity of Sepia officinalis measured using the optomotor response. Canadian Journal of Zoology, 83, 274279.Google Scholar
Guerra, A. (1981). Spatial distribution pattern of Octopus vulgaris. Journal of Zoology (London), 195, 133146.Google Scholar
Guerra, A. & Nixon, M. (1987). Crab and mollusc shell drilling by Octopus vulgaris (Mollusca: Cephalopoda) in the Ria de Vigo (north-west Spain). Journal of Zoology (London), 211, 515523.Google Scholar
Guerra, A., Gonzalez, A. F., Rocha, F. J., Sagarminaga, R. & Canadas, A. (2002a). Planktonic egg masses of the diamond-shaped squid Thysanoteuthis rhombus in the eastern Atlantic and the Mediterranean Sea. Journal of Plankton Research, 24, 333338.Google Scholar
Guerra, A., Gonzalez, A. F., Rocha, F., Segonzac, M. & Gracia, J. (2002b). Observations from submersibles of rare long-arm bathypelagic squids. Sarsia, 87, 189192.Google Scholar
Guerra, A., Hernandez-Urcera, J., Garci, M. E. et al. (2014). Dwellers in dens on sandy bottoms: ecological and behavioural traits of Octopus vulgaris. Scientia Marina, 78, 405414.Google Scholar
Guerra, A., Rocha, F., Gonzalez, A. F. & Gonzalez, J. L. (2006). First observation of sand-covering by the lesser octopus Eledone cirrhosa. Iberus, 24, 2731.Google Scholar
Guibé, M. & Dickel, L. (2011). Embryonic visual experience influences post-hatching shelter preference in cuttlefish. Vie et Milieu – Life and Environment, 61, 243246.Google Scholar
Guibé, M., Boal, J. G. & Dickel, L. (2010). Early exposure to odors changes later visual prey preferences in cuttlefish. Developmental Psychobiology, 52, 833837.Google Scholar
Guibé, M., Poirel, N., Houde, O. & Dickel, L. (2012). Food imprinting and visual generalization in embryos and newly hatched cuttlefish, Sepia officinalis. Animal Behaviour, 84, 213217.Google Scholar
Guilford, T. & Dawkins, M. S. (1991). Receiver psychology and the evolution of animal signals. Animal Behaviour, 42, 114.Google Scholar
Gulland, J. A. & Garcia, S. (1984). Observed patterns in multispecies fisheries, in Exploitation of Marine Communities (ed. May, R. M.), pp. 155190. Berlin: Springer-Verlag.Google Scholar
Gutfreund, Y., Flash, T., Fiorito, G. & Hochner, B. (1998). Patterns of arm muscle activation involved in octopus reaching movements. Journal of Neuroscience, 18, 59765987.Google Scholar
Gutfreund, Y., Flash, T., Yarom, Y. et al. (1996). Organization of octopus arm movements: a model system for studying the control of flexible arms. Journal of Neuroscience, 16, 72977307.Google Scholar
Gutnick, T., Byrne, R. A., Hochner, B. & Kuba, M. (2011). Octopus vulgaris uses visual information to determine the location of its arm. Current Biology, 21, 460462.Google Scholar
Gutsal, D. K. (1989). Underwater observations on distribution and behaviour of cuttlefish Sepia pharaonis in the western Arabian Sea. Biologiya Morya – Marine Biology, 1, 4855.Google Scholar
Haddock, S. H. D., Moline, M. A. & Case, J. F. (2010). Bioluminescence in the sea. Annual Review of Marine Science, 2, 443493.Google Scholar
Hagins, W. A. (1965). Electrical signs of information flow in photoreceptors. Cold Spring Harbor Symposia on Quantitative Biology, 30, 403418.Google Scholar
Hailman, J. P. (1977). Optical Signals. Animal Communication and Light. Bloomington: Indiana Univ. Press.Google Scholar
Hall, D. N. F. (1956). Ink ejection by Cephalopoda. Nature, 7, 663.Google Scholar
Hall, H. (1990). Mugged by a squid! Ocean Realm, 1990, 68.Google Scholar
Hall, K., Fowler, A. J. & Geddes, M. C. (2007). Evidence for multiple year classes of the giant Australian cuttlefish Sepia apama in northern Spencer Gulf, South Australia. Reviews in Fish Biology and Fisheries, 17, 367384.Google Scholar
Hall, K. C. & Hanlon, R. T. (2002). Principal features of the mating system of a large spawning aggregation of the giant Australian cuttlefish Sepia apama (Mollusca: Cephalopoda). Marine Biology, 140, 533545.Google Scholar
Halliday, T. (1983). Information and communication, in Animal Behaviour. 2: Communication (ed. Halliday, T. R. & Slater, P. J. B.), pp. 4381. Oxford: Blackwell Scientific Publications.Google Scholar
Hamabe, M. (1961). Experimental studies on breeding habits and development of the squid, Ommastrephes sloani pacificus Steenstrup I. Copulation. Dobutsugaku Zasshi, 70, 378384.Google Scholar
Hamabe, M. (1964). Study on the migration of squid (Ommastrephes sloani pacificus Steenstrup) as related to the phases of the moon. Bulletin of the Japanese Society of Scientific Fisheries, 30, 209215.Google Scholar
Hamabe, M. & Shimizu, T. (1957). The copulation behavior of Yariika, Loligo bleekeri K. Report on the Japanese Sea Regional Fisheries Research Laboratory, 3, 131136.Google Scholar
Hamasaki, D. I. (1968a). The electroretinogram of the intact anesthetized octopus. Vision Research, 8, 247258.Google Scholar
Hamasaki, D. I. (1968b). The ERG-determined spectral sensitivity of the octopus. Vision Research, 8, 10131021.Google Scholar
Hand, C. (1975). Behaviour of some New Zealand sea anemones and their molluscan and crustacean hosts. New Zealand Journal of Marine & Freshwater Research, 9, 509527.Google Scholar
Hanlon, R. T. (1978). Aspects of the Biology of the Squid Loligo (Doryteuthis) plei in Captivity. Coral Gables: RSMAS, University of Miami.Google Scholar
Hanlon, R. T. (1982). The functional organization of chromatophores and iridescent cells in the body patterning of Loligo plei (Cephalopoda: Myopsida). Malacologia, 23, 89119.Google Scholar
Hanlon, R. T. (1983a). Octopus joubini, in Cephalopod Life Cycles, Vol. I: Species Accounts (ed. Boyle, P. R.), pp. 293310. London: Academic Press.Google Scholar
Hanlon, R. T. (1983b). Octopus briareus, in Cephalopod Life Cycles, Vol. I: Species Accounts (ed. Boyle, P. R.), pp. 251266. London: Academic Press.Google Scholar
Hanlon, R. T. (1988). Behavioral and body patterning characters useful in taxonomy and field identification of cephalopods. Malacologia, 29, 247264.Google Scholar
Hanlon, R. T. (1990). Maintenance, rearing and culture of teuthoid and sepioid squids, in Squid as an Experimental Animal (ed. Gilbert, D. L., Adelman, W. J. & Arnold, J. M.), pp. 3562. New York: Plenum Press.Google Scholar
Hanlon, R. T. (2007). Cephalopod dynamic camouflage. Current Biology, 17, R400R404.Google Scholar
Hanlon, R. T. & Budelmann, B.-U. (1987). Why cephalopods are probably not ‘deaf’. The American Naturalist, 129, 312317.Google Scholar
Hanlon, R. T. & Forsythe, J. W. (1985). Advances in the laboratory culture of octopuses for biomedical research. Laboratory Animal Science, 35, 3340.Google Scholar
Hanlon, R. T. & Forsythe, J. W. (2008). Sexual cannibalism by Octopus cyanea on a Pacific coral reef. Marine and Freshwater Behaviour and Physiology, 41, 1928.Google Scholar
Hanlon, R. T. & Hixon, R. F. (1980). Body patterning and field observations of Octopus burryi Voss, 1950. Bulletin of Marine Science, 30, 749755.Google Scholar
Hanlon, R. T. & Messenger, J. B. (1988). Adaptive coloration in young cuttlefish (Sepia officinalis L.): the morphology and development of body patterns and their relation to behaviour. Philosophical Transactions of the Royal Society of London B, 320, 437487.Google Scholar
Hanlon, R. T. & Shashar, N. (2003). Aspects of the sensory ecology of cephalopods, in Sensory Processing in the Aquatic Environment (ed. Collin, S. P. & Marshall, N. J.), pp. 266282. Heidelberg, Germany: Springer-Verlag.Google Scholar
Hanlon, R. T. & Wolterding, M. R. (1989). Behavior, body patterning, growth and life history of Octopus briareus cultured in the laboratory. American Malacological Bulletin, 7, 2145.Google Scholar
Hanlon, R. T., Ament, S. A. & Gabr, H. (1999). Behavioral aspects of sperm competition in cuttlefish, Sepia officinalis (Sepioidea: Cephalopoda). Marine Biology, 134, 719728.Google Scholar
Hanlon, R. T., Benjamins, S., Beet, A. & Solow, A. (in prep). When camouflage fails: secondary defense tactics of Octopus vulgaris in a coral reef ecosystem. Biological Journal of the Linnean Society.Google Scholar
Hanlon, R. T., Buresch, K., Moustahfid, H. & Staudinger, M. (2013a). Doryteuthis pealeii, longfin inshore squid, in Advances in Squid Biology, Ecology and Fisheries (ed. Rosa, R., O’Dor, R. & Pierce, G. J.), pp. 205240. Hauppauge, New York: Nova Science Publishers, Inc.Google Scholar
Hanlon, R. T., Chiao, C. C., Mäthger, L. M. & Marshall, N. J. (2013b). A fish-eye view of cuttlefish camouflage using in-situ spectrometry. Biological Journal of the Linnean Society, 109, 535551.Google Scholar
Hanlon, R. T., Chiao, C. C., Mäthger, L. M. et al. (2009). Cephalopod dynamic camouflage: bridging the continuum between background matching and disruptive coloration. Philosophical Transactions of the Royal Society B – Biological Sciences, 364, 429437.Google Scholar
Hanlon, R. T., Chiao, C. C., Mäthger, L. M. et al. (2011). Rapid adaptive camouflage in cephalopods, in Animal Camouflage: Mechanisms and Functions (ed. Stevens, M. & Merilaita, S.), pp. 145163. Cambridge: Cambridge University Press.Google Scholar
Hanlon, R. T., Claes, M. F., Ashcraft, S. E. & Dunlap, P. V. (1997). Laboratory culture of the sepiolid squid Euprymna scolopes: a model system for bacteria–animal symbiosis. Biological Bulletin, 192, 364374.Google Scholar
Hanlon, R. T., Conroy, L. A. & Forsythe, J. W. (2008). Mimicry and foraging behaviour of two tropical sand-flat octopus species off North Sulawesi, Indonesia. Biological Journal of the Linnean Society, 93, 2338.Google Scholar
Hanlon, R. T., Cooper, K. M., Budelmann, B. U. & Pappas, T. C. (1990). Physiological color change in squid iridophores. I. Behavior, morphology and pharmacology in Lolliguncula brevis. Cell and Tissue Research, 259, 314.Google Scholar
Hanlon, R. T., Forsythe, J. W. & Boletzky, S. von (1985). Field and laboratory behavior of ‘macrotritopus larvae’ reared to Octopus defilippi Verany, 1851 (Mollusca: Cephalopoda). Vie Milieu, 35, 237242.Google Scholar
Hanlon, R. T., Forsythe, J. W. & Joneschild, D. E. (1999). Crypsis, conspicuousness, mimicry and polyphenism as antipredator defences of foraging octopuses on Indo-Pacific coral reefs, with a method of quantifying crypsis from video tapes. Biological Journal of the Linnean Society, 66, 122.Google Scholar
Hanlon, R. T., Forsythe, J. W. & Messenger, J. B. (1984). Visual discrimination training of laboratory reared octopuses. American Malacological Bulletin, 2, 92.Google Scholar
Hanlon, R. T., Hixon, R. F., Forsythe, J. W. & Hendrix, J. P. (1979). Cephalopods attracted to experimental night lights during a saturation dive at St Croix, U.S. Virgin Islands. Bulletin of the American Malacological Union, 1979, 5358.Google Scholar
Hanlon, R. T., Hixon, R. F. & Hulet, W. H. (1983). Survival, growth, and behavior of the loliginid squids Loligo plei, Loligo pealei, and Lolliguncula brevis (Mollusca: Cephalopoda) in closed sea water systems. Biological Bulletin, 165, 637685.Google Scholar
Hanlon, R. T., Kangas, N. & Forsythe, J. W. (2004). Egg capsule deposition and how behavioral interactions influence spawning rate in the squid Loligo opalescens in Monterey Bay, California. Marine Biology, 145, 923930.Google Scholar
Hanlon, R. T., Maxwell, M. R. & Shashar, N. (1997). Behavioral dynamics that would lead to multiple paternity within egg capsules of the squid Loligo pealei. Biological Bulletin, 193, 212214.Google Scholar
Hanlon, R. T., Maxwell, M. R., Shashar, N., Loew, E. R. & Boyle, K. L. (1999). An ethogram of body patterning behavior in the biomedically and commercially valuable squid Loligo pealei off Cape Cod, Massachusetts. Biological Bulletin, 197, 4962.Google Scholar
Hanlon, R. T., Naud, M. J., Forsythe, J. W. et al. (2007). Adaptable night camouflage by cuttlefish. American Naturalist, 169, 543551.Google Scholar
Hanlon, R. T., Naud, M. J., Shaw, P. W. & Havenhand, J. N. (2005). Behavioural ecology: transient sexual mimicry leads to fertilization. Nature, 430, 212.Google Scholar
Hanlon, R. T., Smale, M. J. & Sauer, W. H. H. (1994). An ethogram of body patterning behaviour in the squid Loligo vulgaris reynaudii on spawning grounds in South Africa. Biological Bulletin, 187, 363372.Google Scholar
Hanlon, R. T., Smale, M. J. & Sauer, W. H. H. (2002). The mating system of the squid Loligo vulgaris reynaudii (Cephalopoda, Mollusca) off South Africa: fighting, guarding, sneaking, mating and egg laying behavior. Bulletin of Marine Science, 71, 331345.Google Scholar
Hanlon, R. T., Turk, P. E., Lee, P. G. & Yang, W. T. (1987). Laboratory rearing of the squid Loligo pealei to the juvenile stage: growth comparisons with fishery data. Fishery Bulletin, 85, 163167.Google Scholar
Hanlon, R. T., Watson, A. C. & Barbosa, A. (2010). A ‘mimic octopus’ in the Atlantic: flatfish mimicry and camouflage by Macrotritopus defilippi. Biological Bulletin, 218, 1524.Google Scholar
Hara, T. & Hara, R. (1972). Cephalopod retinochrome, in Handbook of Sensory Physiology (ed. Dartnall, H. J. A.), pp. 720746. Berlin, Heidelberg, New York: Springer.Google Scholar
Hara, T., Hara, R., Kishigami, A. et al. (1995). Rhodopsin and retinochrome in the retina of a tetrabranchiate cephalopod, Nautilus pompilius. Zoological Science, 12, 195201.Google Scholar
Harden Jones, F. R. (1984). Could fish use inertial clues when on migration?, in Mechanisms of Migration in Fishes (ed. McCleave, J. D., Arnold, G. P., Dodson, J. J. & Neill, W. H.), pp. 6778. New York: Plenum Press.Google Scholar
Hardwick, J. E. (1970). A note on the behavior of the octopod Ocythoe tuberculata. California Fish and Game, 56, 6870.Google Scholar
Hardy, A. C. (1956). The Open Sea. London: Collins.Google Scholar
Harlow, H. F. (1949). The formation of learning sets. Psychological Review, 56, 5165.Google Scholar
Harman, R. F., Young, R. E., Reid, S. B. et al. (1989). Evidence for multiple spawning in the tropical oceanic squid Stenoteuthis oualaniensis (Teuthoidea: Ommastrephidae). Marine Biology, 101, 513519.Google Scholar
Harper, D. G. C. (1991). Communication, in Behavioural Ecology. An Evolutionary Approach (ed. Krebs, J. R. & Davies, N. B.), pp. 374397. Oxford: Blackwell Scientific Publications.Google Scholar
Hartline, P. H., Hurley, A. C. & Lange, G. D. (1979). Eye stabilization by statocyst mediated oculomotor reflex in Nautilus. Journal of Comparative Physiology, 132, 117126.Google Scholar
Hartwick, E. B. (1983). Octopus dofleini, in Cephalopod Life Cycles, Vol. 1: Species Accounts (ed. Boyle, P. R.), pp. 277291. London: Academic Press.Google Scholar
Hartwick, E. B. & Thorarinsson, G. (1978). Den associates of the giant Pacific octopus, Octopus dofleini (Wulker). Ophelia, 17, 163166.Google Scholar
Hartwick, E. B., Ambrose, R. F. & Robinson, S. M. C. (1984). Den utilization and the movements of tagged Octopus dofleini. Marine Behaviour and Physiology, 11, 95110.Google Scholar
Hartwick, E. B., Breen, P. A. & Tulloch, L. (1978). A removal experiment with Octopus dofleini (Wulker). Journal of the Fisheries Research Board of Canada, 35, 14921495.Google Scholar
Hartwick, E. B., Thorarinsson, G. & Tulloch, L. (1978). Antipredator behavior in Octopus dofleini (Wulker). Veliger, 21, 263264.Google Scholar
Harvey, P. H. & Arnold, S. J. (1982). Female mate choice and runaway sexual selection. Nature, 297, 533534.Google Scholar
Harvey, P. H. & Bradbury, J. W. (1991). Sexual selection, in Behavioural Ecology. An Evolutionary Approach (ed. Krebs, J. R. & Davies, N. B.), pp. 203233. Oxford: Blackwell Scientific Publications.Google Scholar
Harvey, P. H. & Pagel, M. D. (1991). The Comparative Method in Evolutionary Biology. Oxford: Oxford University Press.Google Scholar
Haven, N. (1977). The reproductive biology of Nautilus pompilius in the Philippines. Marine Biology, 42, 177184.Google Scholar
Hazlett, B. A. (1977). Quantitative Methods in the Study of Animal Behavior. New York: Academic Press.Google Scholar
Hedge, J., Bart, E. & Kersten, D. (2008). Fragment-based learning of visual object categories. Current Biology, 18, 597601.Google Scholar
Hendrix, J. P. J., Hulet, W. H. & Greenberg, M. J. (1981). Salininity tolerance and the responses to hypoosmotic stress of the bay squid Lolliguncula brevis, an euryhaline cephalopod mollusc. Comparative Biochemistry and Physiology, 69A, 641648.Google Scholar
Hernandez-Urcera, J., Garci, M. E., Roura, A. et al. (2014). Cannibalistic behavior of octopus (Octopus vulgaris) in the wild. Journal of Comparative Psychology, 128, 427430.Google Scholar
Herring, P. J. (1977). Luminescence in cephalopods and fish. Symposia of the Zoological Society of London, 38, 127159.Google Scholar
Herring, P. J. (1988). Luminescent organs, in The Mollusca, Vol. 11: Form and Function (ed. Trueman, E. R. & Clarke, M. R.), pp. 449489. London: Academic Press.Google Scholar
Herring, P. J., Dilly, P. N. & Cope, C. (1985). The photophore morphology of Selenoteuthis scintillans Voss and other lycoteuthids (Cephalopoda: Lycoteuthidae). Journal of Zoology (London), 206, 567589.Google Scholar
Herring, P. J., Dilly, P. N. & Cope, C. (1992). Different types of photophore in the oceanic squids Octopoteuthis and Taningia (Cephalopoda, Octopoteuthidae). Journal of Zoology, 227, 479491.Google Scholar
Heupel, M. R., Semmens, J. M. & Hobday, A. J. (2006). Automated acoustic tracking of aquatic animals: scales, design and deployment of listening station arrays. Marine and Freshwater Research, 57, 113.Google Scholar
Hill, A. V. & Solandt, D. Y. (1935). Myograms from the chromatophores of Sepia. Journal of Physiology (London), 83, 13P14P.Google Scholar
Hinde, R. A. (1970). Animal Behaviour: A Synthesis of Ethology and Comparative Psychology. Kogakusha: McGraw-Hill.Google Scholar
Hirohashi, N., Alvarez, L., Shiba, K. et al. (2013). Sperm from sneaker male squids exhibit chemotactic swarming to CO2. Current Biology, 23, 775781.Google Scholar
Hixon, R. F. (1980). Growth, Reproductive Biology, Distribution and Abundance of Three Species of Loliginid Squid (Myopsida, Cephalopoda) in the Northwest Gulf of Mexico. Ph.D. Dissertation. Coral Gables, FL: University of Miami.Google Scholar
Hixon, R. F., Hanlon, R. T., Gillespie, S. M. & Griffin, W. L. (1980). Squid fishery in Texas: biological, economic, and market considerations. Marine Fisheries Review, 42, 4450.Google Scholar
Hoare, D. J. & Krause, J. (2003). Social organisation, shoal structure and information transfer. Fish and Fisheries, 4, 269279.Google Scholar
Hoare, D. J., Couzin, I. D., Godin, J. G. J. & Krause, J. (2004). Context-dependent group size choice in fish. Animal Behaviour, 67, 155164.Google Scholar
Hochachka, P. W. (1994). Oxygen efficient design of cephalopod muscle metabolism. Marine and Freshwater Behaviour and Physiology, 25, 6167.Google Scholar
Hochberg, F. G. & Couch, J. A. (1971). Biology of cephalopods, in Tektite II, Scientists in the Sea, Mission 8-50, pp. VI-221–VI-228. Washington DC: US Department of the Interior.Google Scholar
Hochner, B. (2012). An embodied view of octopus neurobiology. Current Biology, 22, R887R892.Google Scholar
Hochner, B. (2013). How nervous systems evolve in relation to their embodiment: what we can learn from octopuses and other molluscs. Brain Behavior and Evolution, 82, 1930.Google Scholar
Hochner, B. & Shomrat, T. (2013). The neurophysiological basis of learning and memory in advanced invertebrates: the octopus and the cuttlefish, in Invertebrate Learning and Memory (ed. Menzel, R. & Benjamin, P. R.), pp. 303317. Dusseldorf, Germany: Elsevier/Academic Press.Google Scholar
Hochner, B. & Shomrat, T. (2014). The neurophysiological basis of learning and memory in an advanced invertebrate: the octopus, in Cephalopod Cognition (ed. Darmaillacq, A. S., Dickel, L. & Mather, J.), pp. 7293. Cambridge: Cambridge University Press.Google Scholar
Hochner, B., Brown, E. R., Langella, M., Shomrat, T. & Fiorito, G. (2003). A learning and memory area in the octopus brain manifests a vertebrate-like long-term potentiation. Journal of Neurophysiology, 90, 35473554.Google Scholar
Hochner, B., Shomrat, T. & Fiorito, G. (2006). The octopus: a model for a comparative analysis of the evolution of learning and memory mechanisms. Biological Bulletin, 210, 308317.Google Scholar
Hockett, C. F. (1960). The origin of speech. Scientific American, 203, 8996.Google Scholar
Hodgkin, A. L. (1964). The Conduction of the Nervous Impulse. Liverpool: Liverpool University Press.Google Scholar
Hofmeister, J. K., Alupay, J. S., Ross, R. & Caldwell, R. L. (2011). Observations on mating behavior and development in the lesser Pacific striped octopus, Octopus chierchiae (Jatta, 1889). Integrative and Comparative Biology, 51, E58E58.Google Scholar
Holme, N. A. (1974). The biology of Loligo forbesi Steenstrup (Mollusca: Cephalopoda) in the Plymouth area. Journal of the Marine Biological Association of the United Kingdom, 54, 481503.Google Scholar
Holmes, W. (1940). The colour changes and colour patterns of Sepia officinalis L. Proceedings of the Zoological Society of London A, 110, 235.Google Scholar
Houck, B. A. (1982). Temporal spacing in the activity patterns of three Hawaiian shallow-water octopods. The Nautilus, 96, 152156.Google Scholar
House, M. R. (1988). Major features of cephalopod evolution, in Cephalopods – Present and Past (ed. Wiedmann, J. & Kullmann, J.), pp. 116. Stuttgart: E. Schweizerbart’sche Verlagsbuchhandlung.Google Scholar
Hoving, H. J. T. (2008). Reproductive biology of oceanic decapodiform cephalopods, Ph.D. Thesis, pp. 184. Haren, Netherlands: University of Groningen.Google Scholar
Hoving, H. J. T. & Robison, B. H. (2012). Vampire squid: detritivores in the oxygen minimum zone. Proceedings of the Royal Society B – Biological Sciences, 279, 45594567.Google Scholar
Hoving, H. J. T. & Vecchione, M. (2012). Mating behavior of a deep-sea squid revealed by in situ videography and the study of archived specimens. Biological Bulletin, 223, 263267.Google Scholar
Hoving, H. J. T., Bush, S. L. & Robison, B. H. (2012). A shot in the dark: same-sex sexual behaviour in a deep-sea squid. Biology Letters, 8, 287290.Google Scholar
Hoving, H. J. T., Lipinski, M. R., Videler, J. J. & Bolstad, K. S. R. (2010). Sperm storage and mating in the deep-sea squid Taningia danae Joubin, 1931 (Oegopsida: Octopoteuthidae). Marine Biology, 157, 393400.Google Scholar
Hoving, H. J. T., Zeidberg, L. D., Benfield, M. C. et al. (2013). First in situ observations of the deep-sea squid Grimalditeuthis bonplandi reveal unique use of tentacles. Proceedings of the Royal Society B – Biological Sciences, 280, 20131463.Google Scholar
Hu, M. Y., Yan, H. Y., Chung, W. S., Shiao, J. C. & Hwang, P. P. (2009). Acoustically evoked potentials in two cephalopods inferred using the auditory brainstem response (ABR) approach. Comparative Biochemistry and Physiology A – Molecular & Integrative Physiology, 153, 278283.Google Scholar
Huang, K. L. & Chiao, C. C. (2011). Can cuttlefish learn by observing others? Journal of Shellfish Research, 30, 10081009.Google Scholar
Hubbard, R. & St George, R. C. C. (1958). The rhodopsin system of the squid. Journal General Physiology, 41, 501528.Google Scholar
Hubbard, S. J. (1960). Hearing and the octopus statocyst. Journal of Experimental Biology, 37, 845853.Google Scholar
Huffard, C. L. (2006). Locomotion by Abdopus aculeatus (Cephalopoda: Octopodidae): walking the line between primary and secondary defenses. Journal of Experimental Biology, 209, 36973707.Google Scholar
Huffard, C. L. (2007). Ethogram of Abdopus aculeatus (D’Orbigny, 1834) (Cephalopoda: Octopodidae): can behavioural characters inform octopodid taxonomy and systematics? Journal of Molluscan Studies, 73, 185193.Google Scholar
Huffard, C. L. & Godfrey-Smith, P. (2010). Field observations of mating in Octopus tetricus Gould, 1852 and Amphioctopus marginatus (Taki, 1964) (Cephalopoda: Octopodidae). Molluscan Research, 30, 8186.Google Scholar
Huffard, C. L. & Hochberg, F. G. (2005). Description of a new species of the genus Amphioctopus (Mollusca: Octopodidae) from the Hawai’ian Islands. Molluscan Research, 25, 113128.Google Scholar
Huffard, C. L., Boneka, F. & Full, R. J. (2005). Underwater bipedal locomotion by octopuses in disguise. Science, 307, 1927.Google Scholar
Huffard, C. L., Caldwell, R. L. & Boneka, F. (2008). Mating behavior of Abdopus aculeatus (d’Orbigny 1834) (Cephalopoda: Octopodidae) in the wild. Marine Biology, 154, 353362.Google Scholar
Huffard, C. L., Caldwell, R. L. & Boneka, F. (2010). Male–male and male–female aggression may influence mating associations in wild octopuses (Abdopus aculeatus). Journal of Comparative Psychology, 124, 3846.Google Scholar
Huffard, C. L., Caldwell, R. L., DeLoach, N. et al. (2008). Individually unique body color patterns in octopus (Wunderpus photogenicus) allow for photoidentification. PLoS ONE, 3, e3732.Google Scholar
Huffard, C. L., Saarman, N., Hamilton, H. & Simison, W. B. (2010). The evolution of conspicuous facultative mimicry in octopuses: an example of secondary adaptation? Biological Journal of the Linnean Society, 101, 6877.Google Scholar
Hughes, R. N. (1980). Optimal foraging theory in the marine context. Oceanography and Marine Biology Annual Reviews, 18, 423481.Google Scholar
Hugo, V. (1866). Toilers of the Sea. Paris.Google Scholar
Hulet, W. H., Hanlon, R. T. & Hixon, R. F. (1980). Lolliguncula brevis – a new squid species for the neuroscience laboratory. Trends in Neuroscience, 3, iv–v.Google Scholar
Humphrey, N. K. (1976). The social function of intellect, in Growing Points in Ethology (ed. Bateson, P. P. G. & Hinde, R. A.), pp. 303317. London: Cambridge University Press.Google Scholar
Humphries, D. A. & Driver, P. M. (1970). Protean defence by prey animals. Oecologia, 5, 285302.Google Scholar
Humphries, S., Evans, J. P. & Simmons, L. W. (2008). Sperm competition: linking form to function. BMC Evolutionary Biology, 8, 319.Google Scholar
Hunt, J. C. & Seibel, B. A. (2000). Life history of Gonatus onyx (Cephalopoda: Teuthoidea): ontogenetic changes in habitat, behavior and physiology. Marine Biology, 136, 543552.Google Scholar
Hunt, J. C., Zeidberg, L. D., Hamner, W. M. & Robison, B. H. (2000). The behaviour of Loligo opalescens (Mollusca: Cephalopoda) as observed by a remotely operated vehicle (ROV). Journal of the Marine Biological Association of the United Kingdom, 80, 873883.Google Scholar
Hurley, A. C. (1976). Feeding behavior, food consumption, growth, and respiration of the squid Loligo opalescens raised in the laboratory. Fishery Bulletin, 74, 176182.Google Scholar
Hurley, A. C. (1977). Mating behavior of the squid Loligo opalescens. Marine Behaviour and Physiology, 4, 195203.Google Scholar
Hurley, A. C. (1978). School structure of the squid Loligo opalescens. Fishery Bulletin, 76, 433442.Google Scholar
Hurley, A. C., Lange, G. D. & Hartline, P. H. (1978). Adjustable pinhole camera eye of Nautilus. Journal of Experimental Zoology, 205, 3743.Google Scholar
Hurley, G. V. & Dawe, E. G. (1980). Tagging studies on squid (Illex illecebrosus) in the Newfoundland area. North Atlantic Fisheries Organization Scientific Council Research Document, 80/II/33, #072, 111.Google Scholar
Hvorecny, L. M., Grudowski, J. L., Blakeslee, C. J. et al. (2007). Octopuses (Octopus bimaculoides) and cuttlefishes (Sepia pharaonis, S. officinalis) can conditionally discriminate. Animal Cognition, 10, 449459.Google Scholar
Ibanez, C. M. & Keyl, F. (2010). Cannibalism in cephalopods. Reviews in Fish Biology and Fisheries, 20, 123136.Google Scholar
Ikeda, Y., Sakurai, Y. & Shimazaki, K. (1993). Fertilizing capacity of squid (Todarodes pacificus) spermatozoa collected from various sperm storage sites, with special reference to the role of gelatinous substance from oviducal gland in fertilization and embryonic development. Invertebrate Reproduction & Development, 23, 3944.Google Scholar
Imber, M. J. (1973). The food of grey-faced petrels (Pterodroma macroptera gouldi (Hutton)), with special reference to diurnal migration of their prey. Journal of Animal Ecology, 42, 645662.Google Scholar
Iribarne, O. O. (1990). Use of shelter by the small Patagonian octopus Octopus tehuelchus: availability, selection and effects on fecundity. Marine Ecology Progress Series, 66, 251258.Google Scholar
Iribarne, O. O. (1991). Life history and distribution of the small southwestern Atlantic octopus, Octopus tehuelchus. Journal of Zoology (London), 223, 549565.Google Scholar
Ishii, Y. & Shimada, M. (2010). The effect of learning and search images on predator–prey interactions. Population Ecology, 52, 2735.Google Scholar
Itami, K. (1964). The tagging of Madako (Octopus vulgaris) and its results. Aquaculture, 12, 119125.Google Scholar
Iversen, R. T. B., Perkins, P. J. & Dionne, R. D. (1963). An indication of underwater sound production by squid. Nature, 199, 250251.Google Scholar
Iwata, Y. (2012). Reproductive ecology in loliginid squids. Nippon Suisan Gakkaishi, 78, 665668.Google Scholar
Iwata, Y. & Sakurai, Y. (2007). Threshold dimorphism in ejaculate characteristics in the squid Loligo bleekeri. Marine Ecology Progress Series, 345, 141146.Google Scholar
Iwata, Y., Ito, K. & Sakurai, Y. (2008). Effect of low temperature on mating behavior of squid Loligo bleekeri. Fisheries Science, 74, 13451347.Google Scholar
Iwata, Y., Ito, K. & Sakurai, Y. (2010). Is commercial harvesting of spawning aggregations sustainable? The reproductive status of the squid Loligo bleekeri. Fisheries Research, 102, 286290.Google Scholar
Iwata, Y., Lian, C. L. & Sakurai, Y. (2008). Development of microsatellite markers in the Japanese common squid Todarodes pacificus (Ommastrephidae). Molecular Ecology Resources, 8, 466468.Google Scholar
Iwata, Y., Munehara, H. & Sakurai, Y. (2003). Characterization of microsatellite markers in the squid, Loligo bleekeri (Cephalopoda: Loliginidae). Molecular Ecology Notes, 3, 392393.Google Scholar
Iwata, Y., Munehara, H. & Sakurai, Y. (2005). Dependence of paternity rates on alternative reproductive behaviors in the squid Loligo bleekeri. Marine Ecology Progress Series, 298, 219228.Google Scholar
Iwata, Y., Sakurai, Y. & Shaw, P. (2015). Dimorphic sperm-transfer strategies and alternative mating tactics in loliginid squid. Journal of Molluscan Studies, 81, 147151.Google Scholar
Iwata, Y., Shaw, P., Fujiwara, E., Shiba, K., Kakiuchi, Y. & Hirohashi, N. (2011). Why small males have big sperm: dimorphic squid sperm linked to alternative mating behaviours. BMC Evolutionary Biology, 11, 236.Google Scholar
Izumi, M., Sweeney, A. M., DeMartini, D. et al. (2010). Changes in reflectin protein phosphorylation are associated with dynamic iridescence in squid. Journal of the Royal Society Interface, 7, 549560.Google Scholar
Jackson, G. D. (1994). Application and future potential of statolith increment analysis in squids and sepioids. Canadian Journal of Fisheries and Aquatic Sciences, 51, 26122625.Google Scholar
Jackson, G. D. & Moltschaniwskyj, N. A. (2002). Spatial and temporal variation in growth rates and maturity in the Indo-Pacific squid Sepioteuthis lessoniana (Cephalopoda: Loliginidae). Marine Biology, 140, 747754.Google Scholar
Jackson, G. D., Arkhipkin, A. I., Bizikov, V. A. & Hanlon, R. T. (1993). Laboratory and field corroboration of age and growth from statoliths and gladii of the loliginid squid Sepioteuthis lessoniana (Mollusca: Cephalopoda), in Recent Advances in Cephalopod Fisheries Biology (ed. Okutani, T., O’Dor, R. K. & Kubodera, T.), pp. 189199. Tokyo: Tokai University Press.Google Scholar
Jackson, G. D., Forsythe, J. W., Hixon, R. F. & Hanlon, R. T. (1997). Age, growth, and maturation of Lolliguncula brevis (Cephalopoda: Loliginidae) in the northwestern Gulf of Mexico with a comparison of length-frequency versus statolith age analysis. Canadian Journal of Fisheries and Aquatic Sciences, 54, 29072919.Google Scholar
Jackson, G. D., Lu, C. C. & Dunning, M. (1991). Growth rings within the statolith microstructure of the giant squid Architeuthis. Veliger, 34, 331334.Google Scholar
Jacobson, L. (2005). Essential fish habitat source document: longfin inshore squid, Loligo pealeii, life history and habitat characteristics, 2nd edition. In NOAA Technical Memorandum NMFS-NE-193, pp. 42.Google Scholar
Jander, R., Daumer, K. & Waterman, T. H. (1963). Polarized light orientation by two Hawaiian decapod cephalopods. Zeitschrift fuer vergleichende Physiologie, 46, 383394.Google Scholar
Jantzen, T. M. & Havenhand, J. N. (2003a). Reproductive behavior in the squid Sepioteuthis australis from South Australia: ethogram of reproductive body patterns. Biological Bulletin, 204, 290304.Google Scholar
Jantzen, T. M. & Havenhand, J. N. (2003b). Reproductive behavior in the squid Sepioteuthis australis from South Australia: interactions on the spawning grounds. Biological Bulletin, 204, 305317.Google Scholar
Jatta, G. (1896). I Cefalopodi viventi nel Golfo di Napoli. (Sistematica). Berlin: Verlag Von R. Friedlander & Sohn.Google Scholar
Jefferts, K. (1988). Zoogeography of northeastern Pacific cephalopods, in Cephalopods – Present and Past (ed. Wiedmann, J. & Kullmann, J.), pp. 317339. Stuttgart, Germany: Schweizerbart’sche Verlagsbuchhandlung.Google Scholar
Jereb, P. & Roper, C. F. E. (2010). Cephalopods of the World: An Annotated and Illustrated Catalogue of Cephalopod Species Known to Date, Vol. 2: Myopsid and Oegopsid Squids. Rome, Italy: FAO.Google Scholar
Johnsen, S. (2000). Transparent animals. Scientific American, 282, 8089.Google Scholar
Johnsen, S. (2002). Cryptic and conspicuous coloration in the pelagic environment. Proceedings of the Royal Society B – Biological Sciences, 269, 243256.Google Scholar
Johnsen, S. (2003). Lifting the cloak of invisibility: the effects of changing optical conditions on pelagic crypsis. Integrative and Comparative Biology, 43, 580590.Google Scholar
Johnsen, S., Balser, E. J. & Widder, E. A. (1999). Light-emitting suckers in an octopus. Nature, 398, 113114.Google Scholar
Johnsen, S., Marshall, N. J. & Widder, E. A. (2011). Polarization sensitivity as a contrast enhancer in pelagic predators: lessons from in situ polarization imaging of transparent zooplankton. Philosophical Transactions of the Royal Society B – Biological Sciences, 366, 655670.Google Scholar
Johnson, W. S. & Chase, V. C. (1982). A record of cleaning symbiosis involving Gobiosoma sp. and a large Caribbean octopus. Copeia, 3, 712714.Google Scholar
Johnston, T. D. (1981). Selective costs and benefits in the evolution of learning, in Advances in the Study of Behavior (ed. Rosenblatt, J. S., Hinde, R. A., Beer, C. & Busnel, M. C.). New York: Academic Press.Google Scholar
Johnston, T. D. (1985). Introduction: conceptual issues in the ecological study of learning, in Issues in the Ecological Study of Learning (ed. Johnston, T. D. & Pietrewicz, A. T.). Hillsdale: Erlbaum.Google Scholar
Joll, L. M. (1976). Mating, egg-laying and hatching of Octopus tetricus (Mollusca: Cephalopoda) in the laboratory. Marine Biology, 36, 327333.Google Scholar
Joll, L. M. (1977). The predation of pot-caught western rock lobster (Panulirus longipes cygnus) by octopus. W. Australia Department of Fish Wildlife Reports, 29, 158.Google Scholar
Jolly, A. (1966). Lemur social behavior and primate intelligence. Science, 153, 501506.Google Scholar
Jones, B. W. & Nishiguchi, M. K. (2004). Counterillumination in the Hawaiian bobtail squid, Euprymna scolopes Berry (Mollusca: Cephalopoda). Marine Biology, 144, 11511155.Google Scholar
Jones, E. C. (1963). Tremoctopus violaceus uses Physalia tentacles as weapons. Science, 139, 764766.Google Scholar
Jones, K. A., Jackson, A. L. & Ruxton, G. D. (2011). Prey jitters; protean behaviour in grouped prey. Behavioral Ecology, 22, 831836.Google Scholar
Jordan, M., Chamberlain, J. A. & Chamberlain, R. B. (1988). Response of Nautilus to variation in ambient pressure. Journal of Experimental Biology, 137, 175189.Google Scholar
Josef, N., Amodio, P., Fiorito, G. & Shashar, N. (2012). Camouflaging in a complex environment – octopuses use specific features of their surroundings for background matching. PLoS One, 7, e37579.Google Scholar
Jozet-Alves, C., Bertin, M. & Clayton, N. S. (2013). Evidence of episodic-like memory in cuttlefish. Current Biology, 23, R1033R1035.Google Scholar
Jozet-Alves, C., Moderan, J. & Dickel, L. (2008). Sex differences in spatial cognition in an invertebrate: the cuttlefish. Proceedings of the Royal Society B –Biological Sciences, 275, 20492054.Google Scholar
Jozet-Alves, C., Viblanc, V. A., Romagny, S. et al. (2012). Visual lateralization is task and age dependent in cuttlefish, Sepia officinalis. Animal Behaviour, 83, 13131318.Google Scholar
Kaifu, K., Akamatsu, T. & Segawa, S. (2008). Underwater sound detection by cephalopod statocyst. Fisheries Science, 74, 781786.Google Scholar
Kakinuma, Y., Maki, K., Tsukahara, J. & Tabata, M. (1995). The breeding behavior of Nautilus belauensis. Kagoshima University Research Center for the South Pacific. Occasional Papers, 27, 91105.Google Scholar
Karpov, K. A. & Cailliet, G. M. (1978). Feeding dynamics of Loligo opalescens. California Fish and Game, Fish Bulletin, 169, 4566.Google Scholar
Karson, M. A., Boal, J. G. & Hanlon, R. T. (2003). Experimental evidence for spatial learning in cuttlefish (Sepia officinalis). Journal of Comparative Psychology, 117, 149155.Google Scholar
Kasugai, T. (2000). Reproductive behavior of the pygmy cuttlefish Ideosepius paradoxus in an aquarium. Venus, 59, 3744.Google Scholar
Kasugai, T., Shigeno, S. & Ikeda, Y. (2004). Feeding and external digestion in the Japanese pygmy squid Idiosepius paradoxus (Cephalopoda: Idiosepiidae). Journal of Molluscan Studies, 70, 231236.Google Scholar
Katsanevakis, S. & Verriopoulos, G. (2004). Den ecology of Octopus vulgaris Cuvier, 1797, on soft sediment: availability and types of shelter. Scientia Marina, 68, 147157.Google Scholar
Katz, B. & Miledi, R. (1966). Input–output relations of a single synapse. Nature, 212, 12421245.Google Scholar
Kayes, R. J. (1974). The daily activity pattern of Octopus vulgaris in a natural habitat. Marine Behaviour and Physiology, 2, 337343.Google Scholar
Kear, A. J. (1994). Morphology and function of the mandibular muscles in some coleoid cephalopods. Journal of the Marine Biological Association of the United Kingdom, 74, 801822.Google Scholar
Keenleyside, M. H. A. (1979). Diversity and Adaptation in Fish Behaviour. Berlin: Springer.Google Scholar
Kelman, E. J., Baddeley, R. J., Shohet, A. J. & Osorio, D. (2007). Perception of visual texture and the expression of disruptive camouflage by the cuttlefish, Sepia officinalis. Proceedings of the Royal Society B – Biological Sciences, 274, 13691375.Google Scholar
Kelman, E. J., Osorio, D. & Baddeley, R. J. (2008). A review of cuttlefish camouflage and object recognition and evidence for depth perception. Journal of Experimental Biology, 211, 17571763.Google Scholar
Kelman, E. J., Tiptus, P. & Osorio, D. (2006). Juvenile plaice (Pleuronectes platessa) produce camouflage by flexibly combining two separate patterns. Journal of Experimental Biology, 209, 32883292.Google Scholar
Kemp, D. J., Alcock, J. & Allen, G. R. (2006). Sequential size assessment and multicomponent decision rules mediate aerial wasp contests. Animal Behaviour, 71, 279287.Google Scholar
Kennedy, G. J. A. & Pitcher, T. J. (1975). Experiments on homing in shoals of the European minnow, Phoxinus phoxinus (L.). Transactions of the American Fisheries Society, 104, 452455.Google Scholar
Kennedy, J. S. (1986). Migration, behavioral and ecological, in Migration: Mechanisms and Adaptive Significance (ed. Rankin, M. A.), pp. 526. Port Aransas, TX: University of Texas Marine Science Institute.Google Scholar
Kenyon, K. W. (1975). The Sea Otter in the Eastern Pacific Ocean. New York: Dover.Google Scholar
Kier, W. M. (1985). The musculature of squid arms and tentacles: ultrastructural evidence for functional differences. Journal of Morphology, 185, 223239.Google Scholar
Kier, W. M. (1987). The functional morphology of the tentacle musculature of Nautilus pompilius, in Nautilus. The Biology and Paleobiology of a Living Fossil (ed. Saunders, W. B. & Landman, N. H.), pp. 257269. New York: Plenum.Google Scholar
Kier, W. M. (1988). The arrangement and function of molluscan muscle, in The Mollusca, Vol. 11: Form and Function (ed. Trueman, E. R. & Clarke, M. R.), pp. 211252. San Diego: Academic Press.Google Scholar
Kier, W. M. (1991). Squid cross-striated muscle: the evolution of a specialized muscle fiber type. Bulletin of Marine Science, 49, 389403.Google Scholar
Kier, W. M. & Curtin, N. A. (2002). Fast muscle in squid (Loligo pealei): contractile properties of a specialized muscle fibre type. Journal of Experimental Biology, 205, 19071916.Google Scholar
Kier, W. M. & Smith, A. M. (1990). The morphology and mechanics of octopus suckers. Biological Bulletin, 178, 126136.Google Scholar
Kier, W. M. & Smith, A. M. (2002). The structure and adhesive mechanism of octopus suckers. Integrative and Comparative Biology, 42, 11461153.Google Scholar
Kier, W. M. & Smith, K. K. (1985). Tongues, tentacles and trunks: the biomechanics of movement in muscular-hydrostats. Zoological Journal of the Linneaen Society, 83, 307324.Google Scholar
Kier, W. M. & VanLeeuwen, J. L. (1997). A kinematic analysis of tentacle extension in the squid Loligo pealei. Journal of Experimental Biology, 200, 4153.Google Scholar
Kier, W. M., Messenger, J. B. & Miyan, J. A. (1985). Mechanoreceptors in the fins of the cuttlefish, Sepia officinalis. Journal of Experimental Biology, 119, 369373.Google Scholar
King, A. J. & Adamo, S. A. (2006). The ventilatory, cardiac and behavioural responses of resting cuttlefish (Sepia officinalis L.) to sudden visual stimuli. Journal of Experimental Biology, 209, 11011111.Google Scholar
King, A. J., Adamo, S. A. & Hanlon, R. T. (2003). Squid egg mops provide sensory cues for increased agonistic behaviour between male squid. Animal Behaviour, 66, 4958.Google Scholar
Kito, Y., Seidou, M., Michinomae, M., Partridge, J. C. & Herring, P. J. (1992). Porphyropsin and new deep-sea visual pigment with 4-hydroxyretinal are found in some mesopelagic cephalopods in Atlantic. Zoological Science, 9, 1230.Google Scholar
Kito, Y., Seidou, M., Michinomae, M. & Tokuyama, A. (1987). Photic environment, bioluminescence and vision of a squid Watasenia scintillans. Zoological Science, 4, 1107.Google Scholar
Klages, N. (1989). Food and feeding ecology of emperor penguins in the eastern Weddell Sea. Polar Biology, 9, 385390.Google Scholar
Klages, N. T. W. (1996). Cephalopods as prey. 2. Seals. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 351, 10451052.Google Scholar
Klumpp, D. W. & Nichols, P. D. (1983). A study of food chain in seagrass communities. 2. Food of the rock flathead, Platycephalus laevigatus Cuvier, a major predator in a Posidonia australis seagrass bed. Australian Journal of Marine and Freshwater Research, 34, 745754.Google Scholar
Knight-Jones, E. W. & Morgan, E. (1966). Responses of marine animals to changes in hydrostatic pressure. Oceanography and Marine Biology Annual Reviews, 4, 267299.Google Scholar
Kobayashi, D. R. (1986). Octopus predation on hermit crabs: a test of selectivity. Marine Behaviour and Physiology, 12, 125131.Google Scholar
Kokko, H., Klug, H. & Jennions, M. D. (2012). Unifying cornerstones of sexual selection: operational sex ratio, Bateman gradient and the scope for competitive investment. Ecology Letters, 15, 13401351.Google Scholar
Komak, S., Boal, J. G., Dickel, L. & Budelmann, B. U. (2005). Behavioural responses of juvenile cuttlefish (Sepia officinalis) to local water movements. Marine and Freshwater Behaviour and Physiology, 38, 117125.Google Scholar
Kooyman, G. L., Davis, J. P., Croxall, J. P. & Costa, D. P. (1982). Diving depths and energy requirements of king penguins. Science, 217, 726727.Google Scholar
Koueta, N. & Boucaud-Camou, E. (1989). Etude comparative de la sécrétion des glandes salivaires postérieures des Céphalopodes Decapodes. II. Elaboration de la sécrétion chez les calmars Loligo vulgaris L. et L. forbesi Steenstrup. Bulletin de la Société Zoologique de France, 114, 4754.Google Scholar
Krajewski, J. P., Bonaldo, R. M., Sazima, C. & Sazima, I. (2009). Octopus mimicking its follower reef fish. Journal of Natural History 43, 185190.Google Scholar
Krause, J. (1994). Differential fitness returns in relation to spatial position in groups. Biological Reviews of the Cambridge Philosophical Society, 69, 187206.Google Scholar
Krause, J., Butlin, R. K., Peuhkuri, N. & Pritchard, V. L. (2000). The social organization of fish shoals: a test of the predictive power of laboratory experiments for the field. Biological Reviews of the Cambridge Philosophical Society, 75, 477501.Google Scholar
Krebs, J. R. & Davies, N. B. (1993). An Introduction to Behavioural Ecology. Oxford: Blackwell Scientific Publications.Google Scholar
Krebs, J. R. & Dawkins, R. (1984). Animal rights: mind reading and manipulation, in Behavioural Ecology: An Evolutionary Approach, pp. 380402. Oxford: Blackwell Scientific Publications.Google Scholar
Kristensen, T. K. (1983). Gonatus fabricii, in Cephalopod Life Cycles, Vol. 1 (ed. Boyle, P. R.), pp. 159173. London: Academic Press.Google Scholar
Kröger, B., Vinther, J. & Fuchs, D. (2011). Cephalopod origin and evolution: a congruent picture emerging from fossils, development and molecules. Bioessays, 33, 602613.Google Scholar
Kuba, M., Meisel, D. V., Byrne, R. A., Griebel, U. & Mather, J. A. (2003). Looking at play in Octopus vulgaris, in Coleoid Cephalopods Through Time, Vol. 3 (ed. Warnke, K., Keupp, H. & von Boletzky, S.), pp. 163169: Berliner Palaobiol. Abh.Google Scholar
Kuba, M. J., Byrne, R. A., Meisel, D. V. & Mather, J. A. (2006). When do octopuses play? Effects of repeated testing, object type, age, and food deprivation on object play in Octopus vulgaris. Journal of Comparative Psychology, 120, 184190.Google Scholar
Kubodera, T. & Mori, K. (2005). First-ever observations of a live giant squid in the wild. Proceedings of the Royal Society B: Biological Sciences, 272, 25832586.Google Scholar
Kubodera, T., Koyama, Y. & Mori, K. (2007). Observations of wild hunting behaviour and bioluminescence of a large deep-sea, eight-armed squid, Taningia danae. Proceedings of the Royal Society B – Biological Sciences, 274, 10291034.Google Scholar
Kubodera, T., Piatkowski, U., Okutani, T. & Clarke, M. R. (1998). Taxonomy and zoogeography of the Family Onychoteuthidae (Cephalopoda: Oegopsida). Smithsonian Contributions to Zoology, 586, 277291.Google Scholar
Kuipers, M. R., Pecl, G. T. & Moltschaniwskyj, N. A. (2008). Batch or trickle: understanding the multiple spawning strategy of southern calamary, Sepioteuthis australis (Mollusca: Cephalopoda). Marine & Freshwater Research, 59, 987997.Google Scholar
Kyte, M. A. & Courtney, G. W. (1978). A field observation of aggressive behavior between two North Pacific octopus, Octopus dofleini martini. Veliger, 19, 427428.Google Scholar
Laan, A., Gutnick, T., Kuba, M. J. & Laurent, G. (2014). Behavioral analysis of cuttlefish traveling waves and its implications for neural control. Current Biology, 24, 17371742.Google Scholar
Land, M. F. (1972). The physics and biology of animal reflectors. Progress in Biophysics and Molecular Biology, 24, 75106.Google Scholar
Land, M. F. (1990). Optics of the eyes of marine animals, in Light and Life in the Sea (ed. Herring, P. J., Campbell, A. K., Whitfield, M. & Maddock, L.), pp. 149166. Cambridge: Cambridge University Press.Google Scholar
Land, M. F. (1992). A note on the elongated eye of the octopus Vitreledonella-richardi. Journal of the Marine Biological Association of the United Kingdom, 72, 8992.Google Scholar
Land, M. F. & Nilsson, D.-E. (2012). Animal Eyes. New York: Oxford University Press.Google Scholar
Landman, N. H. & Cochran, J. K. (1987). Growth and longevity of Nautilus, in Nautilus. The Biology and Paleobiology of a Living Fossil (ed. Saunders, W. B. & Landman, N. H.), pp. 401420. New York: Plenum.Google Scholar
Landman, N. H., Cochran, J. K., Cerrato, R. et al. (2004). Habitat and age of the giant squid (Architeuthis sanctipauli) inferred from isotopic analyses. Marine Biology, 144, 685691.Google Scholar
Landman, N. H., Cochran, J. K., Rye, D. M., Tanabe, K. & Arnold, J. M. (1994). Early-life history of Nautilus - evidence from isotopic analyses of aquarium-reared specimens. Paleobiology, 20, 4051.Google Scholar
Landman, N. H., Jones, D. S. & Davis, R. A. (2001). Hatching depth of Nautilus pompilius in Fiji. The Veliger, 44, 333339.Google Scholar
Landry, C., Garant, D., Duchesne, P. & Bernatchez, L. (2001). ‘Good genes as heterozygosity’: the major histocompatibility complex and mate choice in Atlantic salmon (Salmo salar). Proceedings of the Royal Society B – Biological Sciences, 268, 12791285.Google Scholar
Lane, F. W. (1957). Kingdom of the Octopus. The Life History of the Cephalopoda. London: Jarrold.Google Scholar
Langridge, K. V. (2006). Symmetrical crypsis and asymmetrical signalling in the cuttlefish Sepia officinalis. Proceedings of the Royal Society B: Biological Sciences, 273, 959967.Google Scholar
Langridge, K. V. (2009). Cuttlefish use startle displays, but not against large predators. Animal Behaviour, 77, 847856.Google Scholar
Langridge, K. V., Broom, M. & Osorio, D. (2007). Selective signalling by cuttlefish to predators. Current Biology, 17, R1044R1045.Google Scholar
Larcombe, M. F. & Russell, B. C. (1971). Egg laying behaviour of the broad squid Sepioteuthis bilineata (lessoniana). New Zealand Journal of Marine & Freshwater Research, 5, 311.Google Scholar
LaRoe, E. T. (1971). The culture and maintenance of the loliginid squids Sepioteuthis sepioidea and Doryteuthis plei. Marine Biology, 9, 925.Google Scholar
Le Boeuf, B. J., Naito, Y., Huntley, A. C. & Asaga, T. (1989). Prolonged, continuous, deep diving by northern elephant seals. Canadian Journal of Zoology, 67, 25142519.Google Scholar
Lee, H. (1875). The Octopus. The ‘Devil-Fish’ of Fiction and of Fact. London: Chapman and Hall.Google Scholar
Lee, P. G. (1992). Chemotaxis by Octopus maya Voss et Solis in a Y-maze. Journal of Experimental Marine Biology and Ecology, 153, 5367.Google Scholar
Lee, P. G., Turk, P. E., Yang, W. T. & Hanlon, R. T. (1994). Biological characteristics and biomedical applications of the squid Sepioteuthis lessoniana cultured through multiple generations. Biological Bulletin, 186, 328341.Google Scholar
Lee, Y. H., Yan, H. Y. & Chiao, C. C. (2010). Visual contrast modulates maturation of camouflage body patterning in cuttlefish (Sepia pharaonis). Journal of Comparative Psychology, 124, 261270.Google Scholar
Lee, Y.-H., Yan, H. Y. & Chiao, C.-C. (2012). Effects of early visual experience on the background preference in juvenile cuttlefish Sepia pharaonis. Biology Letters, 8, 740743.Google Scholar
Lehmann, U. (1981). The Ammonites: Their Life and Their World. Cambridge: Cambridge University Press.Google Scholar
Lehner, P. N. (1998). Handbook of Ethological Methods 2nd edition. Cambridge: Cambridge University Press.Google Scholar
Leite, T. S. & Mather, J. A. (2008). A new approach to octopuses’ body pattern analysis: a framework for taxonomy and behavioral studies. American Malacological Bulletin, 24, 3141.Google Scholar
Lima, P. A., Nardi, G. & Brown, E. R. (2003). AMPA/kainate and NMDA-like glutamate receptors at the chromatophore neuromuscular junction of the squid: role in synaptic transmission and skin patterning. European Journal of Neuroscience, 17, 507516.Google Scholar
Lima, S. L. & Dill, L. M. (1990). Behavioural decisions made under the risk of predation – a review and prospectus. Canadian Journal of Zoology – Revue Canadienne de Zoologie, 68, 619640.Google Scholar
Lindgren, A. R. (2010). Molecular inference of phylogenetic relationships among Decapodiformes (Mollusca: Cephalopoda) with special focus on the squid Order Oegopsida. Molecular Phylogenetics and Evolution, 56, 7790.Google Scholar
Lindgren, A. R., Giribet, G. & Nishiguchi, M. K. (2004). A combined approach to the phylogeny of Cephalopoda (Mollusca). Cladistics, 20, 454486.Google Scholar
Lindgren, A. R., Pankey, M. S., Hochberg, F. G. & Oakley, T. H. (2012). A multi-gene phylogeny of Cephalopoda supports convergent morphological evolution in association with multiple habitat shifts in the marine environment. BMC Evolutionary Biology, 12, article 129.Google Scholar
Lipinski, M. R. (1985). Laboratory survival of Alloteuthis subulata (Cephalopoda: Loliginidae) from the Plymouth area. Journal of the Marine Biological Association of the United Kingdom, 65, 845855.Google Scholar
Lipinski, M. R. (1987). Food and feeding of Loligo vulgaris reynaudii from St Francis Bay, South Africa. South Africa Journal of Marine Science, 5, 557564.Google Scholar
Lipinski, M. R., Hampton, I., Sauer, W. H. H. & Augustyn, C. J. (1998). Daily net emigration from a spawning concentration of chokka (Loligo vulgaris reynaudii d’Orbigny, 1845) in Kromme Bay, South Africa squid. ICES Journal of Marine Science, 55, 258270.Google Scholar
Lohmann, K. J. & Willows, A. O. D. (1987). Lunar-modulated geomagnetic orientation by a marine mollusk. Science, 235, 331334.Google Scholar
Loi, P. K., Saunders, R. G., Young, D. C. & Tublitz, N. J. (1996). Peptidergic regulation of chromatophore function in the European cuttlefish Sepia officinalis. Journal of Experimental Biology, 199, 11771187.Google Scholar
Long, T. M., Hanlon, R. T., Ter Maat, A. & Pinsker, H. M. (1989). Non-associative learning in the squid Lolliguncula brevis (Mollusca, Cephalopoda). Marine Behaviour and Physiology, 16, 19.Google Scholar
Longley, W. H. (1918). Marine camoufleurs and their camouflage: the present and prospective significance of facts regarding the coloration of tropical fishes. Annual Report of the Smithsonian Institution, 475–485.Google Scholar
Lott, D. F. (1991). Intraspecific Variation in the Social Systems of Wild Vertebrates. Cambridge: Cambridge University Press.Google Scholar
Lu, C. C. & Clarke, M. R. (1975). Vertical distribution of cephalopods at 40° N, 52° N and 60° N at 20° W in the North Atlantic. Journal of the Marine Biological Association of the United Kingdom, 55, 143163.Google Scholar
Lucero, M. T., Farrington, H. & Gilly, W. F. (1994). Quantification of L-DOPA and dopamine in squid ink – implications for chemoreception. Biological Bulletin, 187, 5563.Google Scholar
Lucero, M. T., Horrigan, F. T. & Gilly, W. F. (1992). Electrical responses to chemical stimulation of squid olfactory receptor cells. Journal of Experimental Biology, 162, 231249.Google Scholar
Lum-Kong, A. (1993). Oogenesis, fecundity and pattern of spawning in Loligo forbesi (Cephalopoda: Loliginidae). Malacological Review, 26, 8188.Google Scholar
Lum-Kong, A., Pierce, G. J. & Yau, C. (1992). Timing of spawning and recruitment in Loligo forbesi (Cephalopoda: Loliginidae) in Scottish waters. Journal of the Marine Biological Association of the United Kingdom, 72, 301311.Google Scholar
Lutz, R. A. & Voight, J. R. (1994). Close encounter in the deep. Nature, 371, 563.Google Scholar
Lythgoe, J. N. (1972). The adaptation of visual pigments to the photic environment, in Handbook of Sensory Physiology, Vol. VII/1, Photochemistry of Vision (ed. Dartnall, H. J. A.). Berlin, Heidelberg, New York: Springer.Google Scholar
MacGinitie, G. E. & MacGinitie, N. (1968). Natural History of Marine Animals. New York: McGraw-Hill Book Company.Google Scholar
Machin, K. L. (2005). Avian pain: physiology and evaluation. Compendium on Continuing Education for the Practicing Veterinarian, 27, 98.Google Scholar
Macia, S., Robinson, M. P., Craze, P., Dalton, R. & Thomas, J. D. (2004). New observations on airborne jet propulsion (flight) in squid, with a review of previous reports. Journal of Molluscan Studies, 70, 297299.Google Scholar
Mackintosh, N. J. (1965). Discrimination learning in the Octopus. Animal Behaviour Supplement, 1, 129134.Google Scholar
Mackintosh, N. J. (1974). The Psychology of Animal Learning. London: Academic Press.Google Scholar
Mackintosh, N. J. & Mackintosh, J. (1963). Reversal learning in Octopus, with and without irrelevant cues. Quarterly Journal of Experimental Psychology, 15, 236242.Google Scholar
Mackintosh, N. J. & Mackintosh, J. (1964a). Performance of Octopus over a series of reversals of a simultaneous discrimination. Animal Behaviour, 12, 321324.Google Scholar
Mackintosh, N. J. & Mackintosh, J. (1964b). The effect of overtraining on a nonreversal shift in Octopus. Journal of Genetic Psychology, 106, 373377.Google Scholar
Macy, W. K. (1982). Feeding patterns of the long-finned squid, Loligo pealei, in New England waters. Biological Bulletin, 162, 2838.Google Scholar
Macy, W. K. & Brodziak, J. K. T. (2001). Seasonal maturity and size at age of Loligo pealeii in waters of southern New England. ICES Journal of Marine Science, 58, 852864.Google Scholar
Maddock, L. & Young, J. Z. (1984). Some dimensions of the angular acceleration receptor systems of cephalopods. Journal of the Marine Biological Association of the United Kingdom, 64, 5579.Google Scholar
Maddock, L. & Young, J. Z. (1987). Quantitative differences among the brains of cephalopods. Journal of Zoology (London), 212, 739767.Google Scholar
Madsen, P. T., Wilson, M., Johnson, M. et al. (2007). Clicking for calamari: toothed whales can echolocate squid Loligo pealeii. Aquatic Biology, 1, 141150.Google Scholar
Major, P. F. (1986). Notes on a predator–prey interaction between common dolphins (Delphinus delphis) and short-finned squid (Illex illecebosus) in Lydonia submarine canyon, western North Atlantic Ocean. Journal of Mammalogy, 67, 769770.Google Scholar
Makino, A. & Miyazaki, T. (2010). Topographical distribution of visual cell nuclei in the retina in relation to the habitat of five species of Decapodiformes (Cephalopoda). Journal of Molluscan Studies, 76, 180185.Google Scholar
Maldonado, H. (1964). The control of attack by Octopus. Zeitschrift für vergleichende Physiologie, 47, 656674.Google Scholar
Maldonado, H. (1968). Effect of electroconvulsive shock on memory in Octopus vulgaris. Zeitschrift für vergleichende Physiologie, 59, 2537.Google Scholar
Maldonado, H. (1969). Further investigations on the effect of electroconvulsive shock (ECS) on memory in Octopus vulgaris. Zeitschrift für vergleichende Physiologie, 63, 113118.Google Scholar
Maldonado, H. (1970). The deimatic reaction in the praying mantis Stagmatoptera biocellata. Zeitschrift für vergleichende Physiologie, 68, 6071.Google Scholar
Mangold, K. (1983a). Octopus vulgaris, in Cephalopod Life Cycles, Vol. 1 (ed. Boyle, P. R.), pp. 335364. London: Academic Press.Google Scholar
Mangold, K. (1983b). Food, feeding and growth in cephalopods. Memoirs of the National Museum of Victoria, 44, 8193.Google Scholar
Mangold, K. (1983c). Eledone moschata, in Cephalopod Life Cycles, Vol. 1 (ed. Boyle, P. R.), pp. 475. London: Academic Press.Google Scholar
Mangold, K. (1987). Reproduction, in Cephalopod Life Cycles, Vol. 2: Comparative Reviews (ed. Boyle, P. R.), pp. 157200. London: Academic Press.Google Scholar
Mangold, K. (1989). Traité de Zoologie – Céphalopodes (Tome V, Fascicule 4). Paris: Masson.Google Scholar
Mangold, K. & Boletzky, S. von (1973). New data on reproductive biology and growth of Octopus vulgaris. Marine Biology, 19, 712.Google Scholar
Mangold, K., Young, R. E. & Nixon, M. (1993). Growth versus maturation in cephalopods, in Recent Advances in Cephalopod Fisheries Biology (ed. Okutani, T., O’Dor, R. K. & Kubodera, T.), pp. 697703. Tokyo: Tokai University Press.Google Scholar
Mangold-Wirz, K. (1963). Biologie des Cephalopodes benthiques et nectoniques de la Mer Catalane. Vie Milieu, Suppl. No. 13, 1–285.Google Scholar
Maniwa, Y. (1976). Attraction of bony fish, squid and crab by sound, in Sound Reception in Fish (ed. Schuijf, A. & Hawkins, A. D.), pp. 271283. Amsterdam: Elsevier.Google Scholar
Mann, T. (1984). Spermatophores. Development, Structure, Biochemical Attributes and Role in the Transfer of Spermatozoa. Berlin: Springer.Google Scholar
Mann, T., Martin, A. W. & Thiersch, J. B. (1970). Male reproductive tract, spermatophores and spermatophoric reaction in the giant octopus of the North Pacific, Octopus dofleini martini. Proceedings of the Royal Society of London B: Biological Sciences, 175, 3161.Google Scholar
Marian, J. E. A. R. (2012). A model to explain spermatophore implantation in cephalopods (Mollusca: Cephalopoda) and a discussion on its evolutionary origins and significance. Biological Journal of the Linnean Society, 105, 711726Google Scholar
Markaida, U. (2006). Food and feeding of jumbo squid Dosidicus gigas in the Gulf of California and adjacent waters after the 1997–98 El Nino event. Fisheries Research, 79, 1627.Google Scholar
Markaida, U., Rosenthal, J. J. C. & Gilly, W. F. (2005). Tagging studies on the jumbo squid (Dosidicus gigas) in the Gulf of California, Mexico. Fishery Bulletin, 103, 219226.Google Scholar
Marler, P. & Hamilton, W. J. (1966). Mechanisms of Animal Behavior. New York: Wiley.Google Scholar
Marliave, J. B. (1981). Neustonic feeding in early larvae of Octopus dofleini (Wulker). Veliger, 23, 350351.Google Scholar
Marshall, N. B. (1954). Aspects of Deep Sea Biology. London: Hutchinson.Google Scholar
Marshall, N. J. & Messenger, J. B. (1996). Colour-blind camouflage. Nature, 382, 408409.Google Scholar
Marshall, N. J., Cronin, T. W. & Wehling, M. F. (2011). New directions in biological research on polarized light. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 366, 615616.Google Scholar
Martin, A. W., Catala-Stucki, I. & Ward, P. D. (1978). The growth rate and reproductive behavior of Nautilus macromphalus. Neues Jahrbuch für Geologie und Palaontologie Abhandlungen, 156, 207225.Google Scholar
Martin, P. & Bateson, P. (2007). Measuring Behaviour. An Introductory Guide. Cambridge: Cambridge University Press.Google Scholar
Martin, R. (1965). On the structure and embryonic development of the giant fibre of the squid Loligo vulgaris. Zeitschrift für Zellforschung 67, 7785.Google Scholar
Martin, R. (1977). The giant nerve fibre system of cephalopods. Recent structural findings. Symposia of the Zoological Society of London, 38, 261275.Google Scholar
Martin, S. J., Grimwood, P. D. & Morris, R. G. M. (2000). Synaptic plasticity and memory: an evaluation of the hypothesis. Annual Review of Neuroscience, 23, 649711.Google Scholar
Mather, J. A. (1978). Mating behavior of Octopus joubini Robson. Veliger, 21, 265267.Google Scholar
Mather, J. A. (1980). Social organization and use of space by Octopus joubini in a semi-natural situation. Bulletin of Marine Science, 30, 848857.Google Scholar
Mather, J. A. (1982). Choice and competition: their effects on occupancy of shell homes by Octopus joubini. Marine Behaviour and Physiology, 8, 285293.Google Scholar
Mather, J. A. (1984). Development of behaviour in Octopus joubini Robson, 1929. Vie Milieu, 34, 1720.Google Scholar
Mather, J. A. (1985). Behavioural interactions and activity of captive Eledone moschata: laboratory investigations of a ‘social’ octopus. Animal Behaviour, 33, 11381144.Google Scholar
Mather, J. A. (1986a). A female-dominated feeding hierarchy in juvenile Sepia officinalis in the laboratory. Marine Behaviour and Physiology, 12, 233244.Google Scholar
Mather, J. A. (1986b). Sand digging in Sepia officinalis: assessment of a cephalopod mollusc’s ‘fixed’ behavior pattern. Journal of Comparative Psychology, 100, 315320.Google Scholar
Mather, J. A. (1988). Daytime activity of juvenile Octopus vulgaris in Bermuda. Malacologia, 29, 6976.Google Scholar
Mather, J. A. (1991a). Foraging, feeding, and prey remains in middens of juvenile Octopus vulgaris (Mollusca: Cephalopoda). Journal of Zoology (London), 224, 2739.Google Scholar
Mather, J. A. (1991b). Navigation by spatial memory and use of visual landmarks in octopuses. Journal of Comparative Physiology A, 168, 491497.Google Scholar
Mather, J. A. (1994). Home choice and modification by juvenile Octopus vulgaris (Mollusca, Cephalopoda) – specialized intelligence and tool use. Journal of Zoology, 233, 359368.Google Scholar
Mather, J. A. (1998). How do octopuses use their arms? Journal of Comparative Psychology, 112, 306316.Google Scholar
Mather, J. A. (2008a). To boldly go where no mollusc has gone before: personality, play, thinking, and consciousness in cephalopods. American Malacological Bulletin, 24, 5158.Google Scholar
Mather, J. A. (2008b). Cephalopod consciousness: behavioural evidence. Consciousness and cognition, 17, 3748.Google Scholar
Mather, J. A. (2010). Vigilance and antipredator responses of Caribbean reef squid. Marine and Freshwater Behaviour and Physiology, 43, 357370.Google Scholar
Mather, J. A. & Anderson, R. C. (1993). Personalities of octopuses (Octopus rubescens). Journal of Comparative Psychology, 107, 336340.Google Scholar
Mather, J. A. & Anderson, R. C. (1999). Exploration, play, and habituation in octopuses (Octopus dofleini). Journal of Comparative Psychology, 113, 333338.Google Scholar
Mather, J. A. & Kuba, M. J. (2013). The cephalopod specialties: complex nervous system, learning, and cognition. Canadian Journal of Zoology – Revue Canadienne de Zoologie, 91, 431449.Google Scholar
Mather, J. A. & Mather, D. L. (2004). Apparent movement in a visual display: the ‘passing cloud’ of Octopus cyanea (Mollusca: Cephalopoda). Journal of Zoology, 263, 8994.Google Scholar
Mather, J. A. & O’Dor, R. K. (1984). Spatial organization of schools of the squid Illex illecebrosus. Marine Behaviour and Physiology, 10, 259271.Google Scholar
Mather, J. A. & O’Dor, R. K. (1991). Foraging strategies and predation risk shape the natural history of juvenile Octopus vulgaris. Bulletin of Marine Science, 49, 256269.Google Scholar
Mather, J. A., Anderson, R. C. & Wood, J. B. (2010). Octopus: The Ocean’s Intelligent Invertebrate. Portland, OR: Timber Press, Inc.Google Scholar
Mather, J. A., Griebel, U. & Byrne, R. A. (2010). Squid dances: an ethogram of postures and actions of Sepioteuthis sepioidea squid with a muscular hydrostatic system. Marine and Freshwater Behaviour and Physiology, 43, 4561.Google Scholar
Mäthger, L. M. (2003). The response of squid and fish to changes in the angular distribution of light. Journal of the Marine Biological Association of the United Kingdom, 83, 849856.Google Scholar
Mäthger, L. M. & Denton, E. J. (2001). Reflective properties of iridophores and fluorescent ‘eyespots’ in the loliginid squid Alloteuthis subulata and Loligo vulgaris. Journal of Experimental Biology, 204, 21032118.Google Scholar
Mäthger, L. M. & Hanlon, R. T. (2006). Anatomical basis for camouflaged polarized light communication in squid. Biology Letters, 2, 494496.Google Scholar
Mäthger, L. M. & Hanlon, R. T. (2007). Malleable skin coloration in cephalopods: selective reflectance, transmission and absorbance of light by chromatophores and iridophores. Cell and Tissue Research, 329, 179186.Google Scholar
Mäthger, L. M., Barbosa, A., Miner, S. & Hanlon, R. T. (2006). Color blindness and contrast perception in cuttlefish (Sepia officinalis) determined by a visual sensorimotor assay. Vision Research, 46, 17461753.Google Scholar
Mäthger, L. M., Bell, G. R. R., Kuzirian, A. M., Allen, J. J. & Hanlon, R. T. (2012). How does the blue-ringed octopus (Hapalochlaena lunulata) flash its blue rings? Journal of Experimental Biology, 215, 37523757.Google Scholar
Mäthger, L. M., Chiao, C. C., Barbosa, A. et al. (2007). Disruptive coloration elicited on controlled natural substrates in cuttlefish, Sepia officinalis. Journal of Experimental Biology, 210, 26572666.Google Scholar
Mäthger, L. M., Chiao, C.-C., Barbosa, A. & Hanlon, R. T. (2008). Color matching on natural substrates in cuttlefish, Sepia officinalis. Journal of Comparative Physiology A, 194, 577585.Google Scholar
Mäthger, L. M., Collins, T. F. T. & Lima, P. A. (2004). The role of muscarinic receptors and intracellular Ca2+ in the spectral reflectivity changes of squid iridophores. Journal of Experimental Biology 207, 17591769.Google Scholar
Mäthger, L. M., Denton, E. J., Marshall, N. J. & Hanlon, R. T. (2009). Mechanisms and behavioural functions of structural coloration in cephalopods. Journal of the Royal Society Interface, 6 Suppl. 2, S149S163.Google Scholar
Mäthger, L. M., Roberts, S. B. & Hanlon, R. T. (2010). Evidence for distributed light sensing in the skin of cuttlefish, Sepia officinalis. Biology Letters, 6, 600603.Google Scholar
Mäthger, L. M., Senft, S. L., Gao, M. et al. (2013). Bright white scattering from protein spheres in color changing, flexible cuttlefish skin. Advanced Functional Materials, 23, 39803989.Google Scholar
Mäthger, L. M., Shashar, N. & Hanlon, R. T. (2009). Do cephalopods communicate using polarized light reflections from their skin? J Exp Biol, 212, 21332140.Google Scholar
Matsui, S., Seidou, M., Horiuchi, S., Uchiyama, I. & Kito, Y. (1988). Adaptations of a deep-sea cephalopod to the photic environment. Evidence for three visual pigments. Journal of General Physiology, 92, 5566.Google Scholar
Matteson, R. S., Benoit-Bird, K. J. & Gilly, W. F. (2009). Humboldt squid distribution in three-dimensional space as measured by acoustics in the Gulf of California. Journal of the Acoustical Society of America, 125.Google Scholar
Mattiello, T., Fiore, G., Brown, E. R., d’Ischia, M. & Palumbo, A. (2010). Nitric oxide mediates the glutamate-dependent pathway for neurotransmission in Sepia officinalis chromatophore organs. Journal of Biological Chemistry, 285, 2415424163.Google Scholar
Maturana, H. R. & Sperling, S. (1963). Unidirectional response to angular acceleration recorded from the middle cristal nerve in the statocyst of Octopus vulgaris. Nature, 197, 815816.Google Scholar
Matzner, H., Gutfreund, Y. & Hochner, B. (2000). Neuromuscular system of the flexible arm of the octopus: physiological characterization. Journal of Neurophysiology, 83, 13151328.Google Scholar
Mauris, E. (1989). Colour patterns and body postures related to prey capture in Sepiola affinis (Mollusca: Cephalopoda). Marine Behaviour and Physiology, 14, 189200.Google Scholar
Mauro, A. (1977). Extra-ocular photoreceptors in cephalopods. Symposia of the Zoological Society of London, 38, 287308.Google Scholar
Maxwell, G. (1965). Ring of Bright Water. London: Longmans, Green.Google Scholar
Maxwell, M. R. & Hanlon, R. T. (2000). Female reproductive output in the squid Loligo pealeii: multiple egg clutches and implications for a spawning strategy. Marine Ecology Progress Series, 199, 159170.Google Scholar
Maynard, D. M. (1967). Organization of central ganglia, in Invertebrate Nervous Systems (ed. Wiersma, C. G. A.). Chicago: University Press.Google Scholar
Maynard Smith, J. (1970). The causes of polymorphism. Symposia of the Zoological Society of London, 26, 371383.Google Scholar
Maynard Smith, J. (1974). The theory of games and the evolution of animal conflicts. Journal of Theoretical Biology, 47, 209221.Google Scholar
Maynard Smith, J. (1982). Do animals convey information about their intentions? Journal of Theoretical Biology, 97, 15.Google Scholar
Maynard Smith, J. & Harper, D. (2003). Animal Signals. Oxford: Oxford University Press.Google Scholar
McClean, R. (1983). Gastropod shells: a dynamic resource that helps shape benthic community structure. Journal of Experimental Marine Biology and Ecology, 69, 151174.Google Scholar
McCleneghan, K. & Ames, J. A. (1976). A unique method of prey capture by a sea otter, Enhydra lutris. Journal of Mammalogy, 57, 410412.Google Scholar
McFall-Ngai, M. J. (1990). Crypsis in the pelagic environment. American Zoologist, 30, 175188.Google Scholar
McFarland, D. (1981). The Oxford Companion to Animal Behavior. Oxford: Oxford University Press.Google Scholar
McGowan, J. A. (1954). Observations on the sexual behavior and spawning of the squid, Loligo opalescens, at LaJolla, CA. California Fish and Game, 40, 4754.Google Scholar
Meisel, D. V., Byrne, R. A., Kuba, M. et al. (2006). Contrasting activity patterns of two related octopus species, Octopus macropus and Octopus vulgaris. Journal of Comparative Psychology, 120, 191197.Google Scholar
Meisel, D. V., Byrne, R. A., Mather, J. A. & Kuba, M. (2011). Behavioral sleep in Octopus vulgaris. Vie et Milieu – Life and Environment, 61, 185190.Google Scholar
Mellin, C., Parrott, L., Andrefouet, S. et al. (2012). Multi-scale marine biodiversity patterns inferred efficiently from habitat image processing. Ecological Applications, 22, 792803.Google Scholar
Melo, Y. & Sauer, W. H. H. (2007). Determining the daily spawning cycle of the chokka squid, Loligo reynaudii off the South African coast. Reviews in Fish Biology and Fisheries, 17, 247257.Google Scholar
Melo, Y. C. & Sauer, W. H. H. (1999). Confirmation of serial spawning in the chokka squid Loligo vulgaris reynaudii off the coast of South Africa. Marine Biology, 135, 307313.Google Scholar
Merilaita, S., Schaefer, H. M. & Dimitrova, M. (2013). What is camouflage through distractive markings? Behavioral Ecology, 24, e1271e1272.Google Scholar
Merskey, H. M. & Bogduk, N. (1994). Classification of Chronic Pain. Seattle: IASP Press.Google Scholar
Messenger, J. B. (1963). Behaviour of young Octopus briareus Robson. Nature, 197, 11861187.Google Scholar
Messenger, J. B. (1968). The visual attack of the cuttlefish, Sepia officinalis. Animal Behaviour, 16, 342357.Google Scholar
Messenger, J. B. (1970). Optomotor responses and nystagmus in intact, blinded and statocystless cuttlefish (Sepia officinalis L.). Journal of Experimental Biology, 53, 789796.Google Scholar
Messenger, J. B. (1971). Two-stage recovery of a response in Sepia. Nature, 232, 202203.Google Scholar
Messenger, J. B. (1973a). Learning performance and brain structure: a study in development. Brain Research, 58, 519523.Google Scholar
Messenger, J. B. (1973b). Learning in the cuttlefish, Sepia. Animal Behaviour, 21, 801826.Google Scholar
Messenger, J. B. (1974). Reflecting elements in cephalopod skin and their importance for camouflage. Journal of Zoology (London), 174, 387395.Google Scholar
Messenger, J. B. (1977a). Evidence that Octopus is colour blind. Journal of Experimental Biology, 70, 4955.Google Scholar
Messenger, J. B. (1977b). Prey-capture and learning in the cuttlefish, Sepia. Symposia of the Zoological Society of London, 38, 347376.Google Scholar
Messenger, J. B. (1979a). The eyes and skin of Octopus: compensating for sensory deficiencies. Endeavour, 3, 9298.Google Scholar
Messenger, J. B. (1979b). Nerves, Brain and Behaviour. London: Arnold.Google Scholar
Messenger, J. B. (1979c). The nervous system of Loligo. IV. The peduncle and olfactory lobes. Philosophical Transactions of the Royal Society of London B, 285, 275309.Google Scholar
Messenger, J. B. (1981). Comparative physiology of vision in Molluscs, in Handbook of Sensory Physiology, Vol. VII/6C, Comparative Physiology and Evolution of Vision in Invertebrates (ed. Autrum, H.), pp. 93200. Berlin, Heidelberg, New York: Springer-Verlag.Google Scholar
Messenger, J. B. (1983). Multimodal convergence and the regulation of motor programs in cephalopods, in Multimodal Convergences in Sensory Systems. Fortschritte der Zoologie 28 (ed. Horn, E.), pp. 7798. Stuttgart: Gustav Fischer.Google Scholar
Messenger, J. B. (1991). Photoreception and vision in molluscs, in Evolution of the Eye and Visual System (ed. Cronly-Dillon, J. R. & Gregory, R. L.), pp. 364367. London: Macmillan Press.Google Scholar
Messenger, J. B. (1996). Neurotransmitters of cephalopods. Invertebrate Neuroscience, 2, 95114.Google Scholar
Messenger, J. B. (2001). Cephalopod chromatophores: neurobiology and natural history. Biological Reviews, 76, 473528.Google Scholar
Messenger, J. B. & Sanders, G. D. (1972). Visual preference and two-cue discrimination learning in Octopus. Animal Behaviour, 20, 580585.Google Scholar
Messenger, J. B. & Young, J. Z. (1999). The radular apparatus of cephalopods. Philosophical Transactions of the Royal Society of London B, 354, 161182.Google Scholar
Messenger, J. B., Cornwell, C. & Reed, C. (1997). L-Glutamate and serotonin are endogenous in squid chromatophore nerves. Journal of Experimental Biology, 200, 30433054.Google Scholar
Messenger, J. B., Wilson, A. P. & Hedge, A. (1973). Some evidence for colour-blindness in Octopus. Journal of Experimental Biology, 59, 7794.Google Scholar
Michinomae, M., Masuda, H., Seidou, M. & Kito, Y. (1994). Structural basis for wavelength discrimination in the banked retina of the firefly squid Watasenia scintillans. Journal of Experimental Biology, 193, 112.Google Scholar
Mikami, S. & Okutani, T. (1977). Preliminary observations on maneuvering, feeding, copulating and spawning behaviors of Nautilus macromphalus in captivity. Japanese Journal of Malacology (Venus), 36, 2941.Google Scholar
Milinski, M. & Parker, G. A. (1991). Competition for resources, in Behavioural Ecology. An Evolutionary Approach (ed. Krebs, J. R. & Davies, N. B.), pp. 137168. Oxford: Blackwell Scientific Publications.Google Scholar
Minton, J. W., Walsh, L. S., Lee, P. G. & Forsythe, J. W. (2001). First multi-generation culture of the tropical cuttlefish Sepia pharaonis Ehrenberg, 1831. Aquaculture International, 9, 379392.Google Scholar
Mirow, S. (1972). Skin color in the squids Loligo pealii and Loligo opalescens. II. Iridophores. Zeitschrift für Zellforschung, 125, 176190.Google Scholar
Miske, V. & Kirchhauser, J. (2006). First record of brooding and early life cycle stages in Wunderpus photogenicus Hochberg, Norman and Finn, 2006 (Cephalopoda: Octopodidae). Molluscan Research, 26, 169171.Google Scholar
Miyan, J. & Messenger, J. B. (1995). Intracellular recordings from the chromatophore lobes of Octopus, in Cephalopod Neurobiology (ed. Abbott, N. J., Williamson, R. & Maddock, L.), pp. 415429. Oxford: Oxford University Press.Google Scholar
Mizerez, A., Weaver, J. C., Pedersen, P. B., et al. (2009). Microstructural and biochemical characterization of the nanoporous sucker rings from Dosidicus gigas. Advanced Materials, 20, 16.Google Scholar
Mobley, A. S., Michel, W. C. & Lucero, M. T. (2008). Odorant responsiveness of squid olfactory receptor neurons. Anatomical Record – Advances in Integrative Anatomy and Evolutionary Biology, 291, 763774.Google Scholar
Moiseev, S. I. (1991). Observation of the vertical distribution and behavior of nektonic squids using manned submersibles. Bulletin of Marine Science, 49, 446456.Google Scholar
Moltschaniwskyj, N. A. (1995). Multiple spawning in the tropical squid Photololigo sp.: what is the cost in somatic growth? Marine Biology, 124, 127135.Google Scholar
Moltschaniwskyj, N. A. & Pecl, G. T. (2007). Spawning aggregations of squid (Sepioteuthis australis) populations: a continuum of ‘microcohorts’. Reviews in Fish Biology and Fisheries, 17, 183195.Google Scholar
Moltschaniwskyj, N. A. & Steer, M. A. (2004). Spatial and seasonal variation in reproductive characteristics and spawning of southern calamary (Sepioteuthis australis): spreading the mortality risk. ICES Journal of Marine Science, 61, 921927.Google Scholar
Moody, M. F. & Parriss, J. R. (1960). The visual system of Octopus. Discrimination of polarized light by Octopus. Nature, 186, 839840.Google Scholar
Moody, M. F. & Parriss, J. R. (1961). The discrimination of polarized light by Octopus: a behavioural and morphological study. Zeitschrift für vergleichende Physiologie, 44, 268291.Google Scholar
Mooney, T. A., Hanlon, R. T., Christensen-Dalsgaard, J. et al. (2010). Sound detection by the longfin squid (Loligo pealeii) studied with auditory evoked potentials: sensitivity to low-frequency particle motion and not pressure. Journal of Experimental Biology, 213, 37483759.Google Scholar
Morejohn, G. V., Harvey, J. T. & Krasnow, L. T. (1978). The importance of Loligo opalescens in the food web of marine vertebrates in Monterey Bay, California, in Department of Fish and Game, Fish Bulletin 169 (ed. Recksiek, C. W. & Frey, H. W.), pp. 6798. Los Angeles: Department of Fish and Game.Google Scholar
Moroshita, T. (1974). Participation in digestion by the proteolytic enzymes of the posterior salivary gland in Octopus – III. Some properties of purified enzymes from the posterior salivary gland. Bulletin of the Japanese Society of Scientific Fisheries, 40, 927936.Google Scholar
Moynihan, M. (1975). Conservatism of displays and comparable stereotyped patterns among cephalopods, in Function and Evolution in Behaviour. Essays in Honor of Professor Niko Tinbergen, F.R.S. (ed. Baerends, G., Beer, C. & Manning, A.), pp. 276291. New York: Oxford University Press.Google Scholar
Moynihan, M. (1983a). Notes on the behavior of Euprymna scolopes (Cephalopoda: Sepiolidae). Behaviour, 85, 2541.Google Scholar
Moynihan, M. (1983b). Notes on the behavior of Idiosepius pygmaeus (Cephalopoda; Idiosepiidae). Behaviour, 85, 4257.Google Scholar
Moynihan, M. (1985a). Communication and Noncommunication by Cephalopods. Bloomington: Indiana University Press.Google Scholar
Moynihan, M. (1985b). Why are cephalopods deaf? The American Naturalist, 125, 465469.Google Scholar
Moynihan, M. & Rodaniche, A. F. (1982). The behavior and natural history of the Caribbean reef squid Sepioteuthis sepioidea. With a consideration of social, signal and defensive patterns for difficult and dangerous environments. Advances in Ethology, 25, 1150.Google Scholar
Moynihan, M. & Rodaniche, F. (1977). Communication, crypsis, and mimicry among cephalopods, in How Animals Communicate (ed. Sebeok, T. A.), pp. 293302. Bloomington: Indiana University Press.Google Scholar
Muñoz, J. L. P., Patino, M. A. L., Hermosilla, C. et al.(2011). Melatonin in octopus (Octopus vulgaris): tissue distribution, daily changes and relation with serotonin and its acid metabolite. Journal of Comparative Physiology A –Neuroethology Sensory Neural and Behavioral Physiology, 197, 789797.Google Scholar
Muntz, W. R. A. (1961). Interocular transfer in Octopus vulgaris. Journal of Comparative and Physiological Psychology, 54, 4955.Google Scholar
Muntz, W. R. A. (1962). Stimulus generalisation following monocular training in Octopus. Journal of Comparative and Physiological Psychology, 55, 535540.Google Scholar
Muntz, W. R. A. (1976). On yellow lenses in mesopelagic animals. Journal of the Marine Biological Association of the United Kingdom, 56, 963976.Google Scholar
Muntz, W. R. A. (1977). Pupillary response of cephalopods. Symposia of the Zoological Society of London, 38, 277285.Google Scholar
Muntz, W. R. A. (1986). The spectral sensitivity of Nautilus pompilius. Journal of Experimental Biology, 126, 513517.Google Scholar
Muntz, W. R. A. (1987). Visual behavior and visual sensitivity of Nautilus pompilius, in Nautilus: The Biology and Paleobiology of a Living Fossil (ed. Saunders, W. B. & Landman, N. H.), pp. 231244. New York: Plenum Press.Google Scholar
Muntz, W. R. A. (1991). Anatomical and behavioural studies on vision in Nautilus and Octopus. American Malacological Bulletin, 9, 6974.Google Scholar
Muntz, W. R. A. (1994a). Spatial summation in the phototactic behavior of Nautilus pompilius. Marine Behaviour and Physiology, 24, 183187.Google Scholar
Muntz, W. R. A. (1994b). Effects of light on the efficacy of traps for Nautilus pompilius. Marine Behaviour and Physiology, 24, 189193.Google Scholar
Muntz, W. R. A. & Gwyther, J. (1988a). Visual discrimination of distance by octopuses. Journal of Experimental Biology, 140, 345353.Google Scholar
Muntz, W. R. A. & Gwyther, J. (1988b). Visual acuity in Octopus pallidus and Octopus australis. Journal of Experimental Biology 134, 119129.Google Scholar
Muntz, W. R. A. & Johnson, M. S. (1978). Rhodopsins of oceanic decapods. Vision Research, 18, 601602.Google Scholar
Muntz, W. R. A. & Raj, U. (1984). On the visual system of Nautilus pompilius. Journal of Experimental Biology, 109, 253263.Google Scholar
Muntz, W. R. A. & Wentworth, S. L. (1987). An anatomical study of the retina of Nautilus pompilius. Biological Bulletin, 173, 387397.Google Scholar
Muntz, W. R. A. & Wentworth, S. L. (1995). Structure of the adhesive surface of the digital tentacles of Nautilus pompilius. Journal of the Marine Biological Association of the United Kingdom, 75, 747750.Google Scholar
Munz, F. W. (1958). Photosensitive pigments from the retinae of certain deep-sea fishes. Journal of Physiology (London), 140, 220225.Google Scholar
Munz, F. W. (1964). The visual pigments of epipelagic and rocky shore fishes. Vision Research, 4, 441454.Google Scholar
Muramatsu, K., Yamamoto, J., Abe, T. et al. (2013). Oceanic squid do fly. Marine Biology, 160, 11711175.Google Scholar
Murata, M. (1990). Oceanic resources of squids. Marine Behaviour and Physiology B, 18, 1971.Google Scholar
Murata, M., Ishii, M. & Osako, M. (1982). Some information on copulation of the oceanic squid Onychoteuthis borealijaponica Okada. Bulletin of the Japanese Society of Scientific Fisheries, 48, 351354.Google Scholar
Myrberg, A. A. (1973). Underwater television – a tool for the marine biologist. Bulletin of Marine Science, 23, 824836.Google Scholar
Nabhitabhata, J. (1998). Distinctive behaviour of Thai pygmy squid, Idiosepius thailandicus Chotiyaputta, Okutani & Chaitiamvong, 1991. Phuket Marine Biological Center Special Publication, 18, 2540.Google Scholar
Nabhitabhata, J. & Suwanamala, J. (2008). Reproductive behaviour and cross-mating of two closely related pygmy squids Idiosepius biserialis and Idiosepius thailandicus (Cephalopoda: Idiosepiidae). Journal of the Marine Biological Association of the United Kingdom, 88, 987993.Google Scholar
Naef, A. (1923). Die Cephalopoden. Systematik. Fauna e Flora del Golfo di Napoli, 35, 1863.Google Scholar
Naef, A. (1928). Die Cephalopoden. Embryologie. Fauna e Flora del Golfo di Napoli, 35, 1357.Google Scholar
Nagasawa, K., Takayanagi, S. & Takami, T. (1993). Cephalopod tagging and marking in Japan: a review, in Recent Advances in Cephalopod Fisherìes Biology (ed. Okutani, T., O’Dor, R. K. & Kubodera, T.), pp. 313329. Tokyo: Tokai University Press.Google Scholar
Nakamura, Y. (1991). Tracking of the mature female of flying squid, Ommastrephes bartrami, by an ultrasonic transmitter. Bulletin of the Hokkaido National Fisheries Research Institute, 55, 205207.Google Scholar
Narvarte, M., Gonzalez, R. A., Storero, L. & Fernandez, M. (2013). Effects of competition and egg predation on shelter use by Octopus tehuelchus females. Marine Ecology Progress Series, 482, 141151.Google Scholar
Natsukari, Y. (1970). Egg-laying behavior, embryonic development and hatched larva of the pygmy cuttlefish, Idiosepius pygmaeus paradoxus Ortmann. Bulletin of the Faculty of Fisheries of Nagasaki University, 30, 1529.Google Scholar
Natsukari, Y. & Tashiro, M. (1991). Neritic squid resources and cuttlefish resources in Japan. Marine Behaviour and Physiology B, 18, 149226.Google Scholar
Naud, M. J. & Havenhand, J. N. (2006). Sperm motility and longevity in the giant cuttlefish, Sepia apama (Mollusca: Cephalopoda). Marine Biology, 148, 559566.Google Scholar
Naud, M. J., Hanlon, R. T., Hall, K. C., Shaw, P. W. & Havenhand, J. N. (2004). Behavioral and genetic assessment of mating success in a natural spawning aggregation of the giant cuttlefish (Sepia apama) in southern Australia. Animal Behaviour, 67, 10431050.Google Scholar
Naud, M. J., Shaw, P. W., Hanlon, R. T. & Havenhand, J. N. (2005). Evidence for biased use of sperm sources in wild female giant cuttlefish (Sepia apama). Proceedings of the Royal Society B: Biological Sciences, 272, 10471051.Google Scholar
Neill, S. R. S. J. (1971). Notes on squid and cuttlefish; keeping, handling and colour-patterns. Pubblicazioni della Stazione Zoologica di Napoli, 39, 6469.Google Scholar
Neill, S. R. S. J. & Cullen, J. M. (1974). Experiments on whether schooling by their prey affects the hunting behaviour of cephalopods and fish predators. Journal of Zoology (London), 172, 549569.Google Scholar
Nesher, N., Levy, G., Grasso, F. W. & Hochner, B. (2014). Self-recognition mechanism between skin and suckers prevents octopus arms from intefering with each other. Current Biology, 24, 12711275.Google Scholar
Nesis, K. N. (1965). Distribution and feeding of young squids Gonatus fabricii (Licht.) in the Labrador Sea and the Norwegian Sea. Oceanology (Washington), 5, 102108.Google Scholar
Nesis, K. N. (1970). The biology of the giant squid of Peru and Chile, Dosidicus gigas. Oceanology (Washington), 10, 108118.Google Scholar
Nesis, K. N. (1982). Cephalopods of the World. Squids, Cuttlefishes, Octopuses, and Allies, translated from Russian by B. F. Levitov (ed. Burgess, L.A.) Neptune City, NJ: TFH Publications.Google Scholar
Nesis, K. N. (1983). Dosidicus gigas, in Cephalopod Life Cycles, Vol. 1 (ed. Boyle, P. R.), pp. 475. London: Academic Press.Google Scholar
Nesis, K. N. (1995). Mating, spawning and death in oceanic cephalopods: a review. Ruthenica, 6, 2364.Google Scholar
Nesis, K. N. & Nikitina, I. V. (1981). Macrotritopus, a planktonic larva of the benthic octopus, Octopus defilippi: identification and distribution. Zoologicheskifi Zhurnal, 60, 835847.Google Scholar
Neumeister, H. & Budelmann, B. U. (1997). Structure and function of the Nautilus statocyst. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 352, 15651588.Google Scholar
Nicol, S. & O’Dor, R. K. (1985). Predatory behaviour of squid (Illex illecebrosus) feeding on surface swarms of euphausiids. Canadian Journal of Zoology, 63, 1517.Google Scholar
Nigmatullin, C. M. & Ostapenko, A. A. (1976). Feeding of Octopus vulgaris Lam. from the northwest African coast. Shellfish & Benthos Committee. International Council for the Exploration of the Sea, C.M. 1976/K, 6, 115.Google Scholar
Nigmatullin, C. M., Nesis, K. N. & Arkhipkin, A. I. (2001). A review of the biology of the jumbo squid Dosidicus gigas (Cephalopoda: Ommastrephidae). Fisheries Research, 54, 919.Google Scholar
Nilsson, D. E., Warrant, E. J., Johnsen, S., Hanlon, R. & Shashar, N. (2012). A unique advantage for giant eyes in giant squid. Current Biology, 22, 683688.Google Scholar
Nilsson, D.-E., Warrant, E. J., Johnsen, S., Hanlon, R. T. & Shashar, N. (2013). The giant eyes of giant squid are indeed unexpectedly large, but not if used for spotting sperm whales. BMC Evolutionary Biology, 13, Article 187.Google Scholar
Nishimura, M. (1961). Frequency characteristics of sea noise and fish sound. Technical Report of Fishing Boat, Tokyo, 15, 111118.Google Scholar
Nixon, M. (1979). Hole-boring in shells by Octopus vulgaris Cuvier in the Mediterranean. Malacologia, 18, 431443.Google Scholar
Nixon, M. (1980). The salivary papilla of Octopus as an accessory radula for drilling shells. Journal of Zoology (London), 190, 5357.Google Scholar
Nixon, M. (1984). Is there external digestion by Octopus? Journal of Zoology (London), 202, 441447.Google Scholar
Nixon, M. (1987). Cephalopod diets, in Cephalopod Life Cycles, Vol. 2: Comparative Reviews (ed. Boyle, P. R.), pp. 201219. London: Academic Press.Google Scholar
Nixon, M. & Budelmann, B. U. (1984). Scale-worms – occasional food of Octopus. Journal of Molluscan Studies, 50, 3942.Google Scholar
Nixon, M. & Dilly, P. N. (1977). Sucker surfaces and prey capture. Symposia of the Zoological Society of London, 38, 447511.Google Scholar
Nixon, M. & Maconnachie, E. (1988). Drilling by Octopus vulgaris (Mollusca: Cephalopoda) in the Mediterranean. Journal of Zoology (London), 216, 687716.Google Scholar
Nixon, M. & Messenger, J. B. (1977). The Biology of Cephalopods. London: Academic Press.Google Scholar
Nixon, M. & Young, J. Z. (2003). The Brains and Lives of Cephalopods. Oxford: Clarendon.Google Scholar
Nixon, M., Maconnachie, E. & Howell, P. G. T. (1980). The effects on shells of drilling by Octopus. Journal of Zoology (London), 191, 7588.Google Scholar
Noakes, D. L. G. & Baylis, J. R. (1990). Behavior, in Methods for Fish Biology (ed. Schreck, C. B. & Moyle, P. B.), pp. 555583. Bethesda, MD: American Fisheries Society.Google Scholar
Norman, M. D. (1991). Octopus cyanea Gray, 1849 (Mollusca, Cephalopoda) in Australian waters – description, distribution and taxonomy. Bulletin of Marine Science, 49, 2038.Google Scholar
Norman, M. D. (1992a). Ameloctopus litoralis, gen. et sp. nov. (Cephalopoda: Octopodidae), a new shallow-water octopus from tropical Australian waters. Invertebrate Taxonomy, 6, 567582.Google Scholar
Norman, M. D. (1992b). Ocellate octopuses (Cephalopoda: Octopodidae) of the Great Barrier Reef, Australia: description of two new species and redescription of Octopus polyzenia Gray, 1949. Memoirs of the Museum of Victoria, 53, 309344.Google Scholar
Norman, M. D. (2000). Cephalopods: A World Guide. Hackenheim: ConchBooks.Google Scholar
Norman, M. D. & Lu, C. C. (1997). Sex in giant squid. Nature, 389, 683684.Google Scholar
Norman, M. D., Finn, J. & Tregenza, T. (1999). Female impersonation as an alternative reproductive strategy in giant cuttlefish. Proceedings of the Royal Society B: Biological Sciences, 266, 13471349.Google Scholar
Norman, M. D., Finn, J. & Tregenza, T. (2001). Dynamic mimicry in an Indo-Malayan octopus. Proceedings of the Royal Society of London B: Biological Sciences, 268, 17551758.Google Scholar
Norman, M. D., Paul, D., Finn, J. & Tregenza, T. (2002). First encounter with a live male blanket octopus: the world’s most sexually size-dimorphic large animal. New Zealand Journal of Marine and Freshwater Research, 36, 733736.Google Scholar
Norris, K. S. & Mohl, B. (1983). Can odontocetes debilitate prey with sound? The American Naturalist, 122, 85104.Google Scholar
Novicki, A., Messenger, J. B., Budelmann, B. U., Terrell, M. L. & Kadekaro, M. (1992). [14C] Deoxyglucose labelling of functional activity in the cephalopod central nervous system. Proceedings of the Royal Society of London B, 249, 7782.Google Scholar
Nyholm, S. V. & Nishiguchi, M. K. (2008). The evolutionary ecology of a sepiolid squid–Vibrio association: from cell to environment. Vie et Milieu – Life and Environment, 58, 175184.Google Scholar
O’Brien, W. J., Browman, H. I. & Evans, B. I. (1990). Search strategies of foraging animals. American Scientist, 78, 152160.Google Scholar
O’Dor, R. (2002). Telemetered cephalopod energetics: swimming, soaring, and blimping. Integrative and Comparative Biology, 42, 10651070.Google Scholar
O’Dor, R., Balch, N. E. & Amaratunga, T. (1982). Laboratory observations of midwater spawning by Illex illecebrosus. NAFO Scientific Council Studies, 9, 69133.Google Scholar
O’Dor, R., Stewart, J., Gilly, W. et al. (2013). Squid rocket science: how squid launch into air. Deep-Sea Research Part II: Tropical Studies in Oceanography, 95, 113118.Google Scholar
O’Dor, R. K. (1983). Illex illecebrosus, in Cephalopod Life Cycles, Vol. 1 (ed. Boyle, P. R.), pp. 175199. London: Academic Press.Google Scholar
O’Dor, R. K. (1988). The energetic limits on squid distributions. Malacologia, 29, 113119.Google Scholar
O’Dor, R. K. (1998). Can understanding squid life-history strategies and recruitment improve management? South African Journal of Marine Science – Suid-Afrikaanse Tydskrif vir Seewetenskap, 20, 193206.Google Scholar
O’Dor, R. K. (2012). The incredible flying squid. New Scientist, 214, 3941.Google Scholar
O’Dor, R. K. & Balch, N. (1985). Properties of Illex illecebrosus egg masses potentially influencing larval oceanographic distribution. NAFO Scientific Council Studies, 9, 6976.Google Scholar
O’Dor, R. K. & Macalaster, E. G. (1983). Bathypolypus arcticus, in Cephalopod Life Cycles, Vol. 1 (ed. Boyle, P. R.), pp. 475. London: Academic Press.Google Scholar
O’Dor, R. K. & Webber, D. M. (1986). The constraints on cephalopods: why squid aren't fish. Canadian Journal of Zoology, 64, 15911605.Google Scholar
O’Dor, R. K. & Webber, D. M. (1991). Invertebrate athletes: trade-offs between transport efficiency and power density in cephalopod evolution. Journal of Experimental Biology, 160, 93112.Google Scholar
O’Dor, R. K. & Wells, M. J. (1978). Reproduction versus somatic growth: hormonal control in Octopus vulgaris. Journal of Experimental Biology, 77, 1531.Google Scholar
O’Dor, R. K. & Wells, M. J. (1987). Energy and nutrient flow, in Cephalopod Life Cycles, Vol. 2: Comparative Reviews (ed. Boyle, P. R.), pp. 109133. London: Academic Press.Google Scholar
O’Dor, R. K., Adamo, S., Aitken, J. P. et al. (2002). Currents as environmental constraints on the behavior, energetics and distribution of squid and cuttlefish. Bulletin of Marine Science, 71, 601617.Google Scholar
O’Dor, R. K., Andrade, Y., Webber, D. M. et al. (1998). Applications and performance of Radio-Acoustic Positioning and Telemetry (RAPT) systems. Hydrobiologia, 372, 18.Google Scholar
O’Dor, R. K., Carey, F. G., Webber, D. M. & Voegeli, F. M. (1991). Behavior and energenetics of Azorean squid, Loligo forbesi, in Biotelemetry XI, Proceedings of the Eleventh International Symposium on Biotelemetry (ed. Uchiyama, A. & Amlaner, C. J. Jr), pp. 191195. Tokyo: Waseda University Press.Google Scholar
O’Dor, R. K., Forsythe, J., Webber, D. M., Wells, J. & Wells, M. J. (1993). Activity levels of Nautilus. Nature, 362, 626627.Google Scholar
O’Dor, R. K., Helm, P. & Balch, N. (1985). Can rhynchoteuthions suspension feed? (Mollusca: Cephalopoda). Vie et Milieu, 35, 267271.Google Scholar
O’Dor, R. K., Hoar, J. A., Webber, D. M. et al. (1995). Squid (Loligo forbesi) performance and metabolic rates in nature. Marine and Freshwater Behaviour and Physiology, 25, 163177.Google Scholar
O’Dor, R. K., Stewart, J., Gilly, W. et al. (2013). Squid rocket science: how squid launch into air. Deep-Sea Research Part II: Tropical Studies in Oceanography, 95, 113118.Google Scholar
O’Dor, R. K., Vessey, E. & Amaratunga, T. (1980). Factors affecting fecundity and larval distribution in the squid, Illex illecebrosus. North Atlantic Fisheries Organization Scientific Council Research Document, 2, 19.Google Scholar
O’Dor, R. K., Wells, J. & Wells, M. J. (1990). Speed, jet pressure and oxygen consumption relationships in free-swimming Nautilus. Journal of Experimental Biology, 154, 383396.Google Scholar
Okutani, T. (1960). Argonauta boettgeri preys on Cavolinia tridentata. Venus, 21, 3941.Google Scholar
Okutani, T. (1974). Epipelagic decapod cephalopods collected by micronekton tows during the EASTROPAC Expeditions, 1967–1968 (Systematic Part). Bulletin of the Tokai Regional Fisheries Research Laboratory, 80, 29118.Google Scholar
Okutani, T. (1983). Todarodes pacificus, in Cephalopod Life Cycles, Vol. 1 (ed. Boyle, P. R.), pp. 475. London: Academic Press.Google Scholar
Okutani, T. & Osuga, K. (1986). A peculiar nesting behavior of Ocythoe tuberculata in the test of a gigantic salp, Tethys vagina. Venus: Japanese Journal of Malacology, 45, 6769.Google Scholar
Oliveira, R. F., Taborsky, M. & Brockmann, H. J. (ed.) (2008). Alternative Reproductive Tactics: An Integrative Approach. Cambridge: Cambridge University Press.Google Scholar
Olofsson, M., Eriksson, S., Jakobsson, S. & Wiklund, C. (2012). Deimatic display in the European swallowtail butterfly as a secondary defence against attacks from great tits. PLoS One, 7.Google Scholar
Ord, T. J., Peters, R. A., Evans, C. S. & Taylor, A. J. (2002). Digital video playback and visual communication in lizards. Animal Behaviour, 63, 879890.Google Scholar
Orelli, M. v. (1962). Die Ubertragung der spermatophore von Octopus vulgaris and Eledone (Cephalopoda). Revue Suisse de Zoologie, 69, 193202.Google Scholar
Orlov, O. Y. & Byzov, A. L. (1961). Colorimetric research on the vision of molluscs (Cephalopoda). Doklady Akademii NaukSSSR, 139, 723725.Google Scholar
Orlov, O. Y. & Byzov, A. L. (1962). Vision in cephalopod molluscs. Priroda Moskva, 3, 115118.Google Scholar
Ormond, R. F. G. (1980). Aggressive mimicry and other interspecific feeding associations among Red Sea coral reef predators. Journal of Zoology, 191, 247262.Google Scholar
O’Shea, S. & Bolstad, K. S. (2004). First records of egg masses of Nototodarus gouldi McCoy, 1888 (Mollusca: Cephalopoda: Ommastrephidae), with comments on egg-mass susceptibility to damage by fisheries trawl. New Zealand Journal of Zoology, 31, 161166.Google Scholar
Otis, T. S. & Gilly, W. F. (1990). Jet-propelled escape in the squid Loligo opalescens: concerted control by giant and non-giant motor axon pathways. Proceedings of the National Academy of Sciences USA, 87, 29112915.Google Scholar
Packard, A. (1961). Sucker display of Octopus. Nature, 190, 736737.Google Scholar
Packard, A. (1963). The behaviour of Octopus vulgaris. Bulletin de l’Institut Océanographique (Monaco), No. 1 D, 35–49.Google Scholar
Packard, A. (1969a). Jet propulsion and giant fibre response of Loligo. Nature, 221, 875.Google Scholar
Packard, A. (1969b). Visual acuity and eye growth in Octopus vulgaris (Lamarck). Monitore Zoologico Italiano, 3, 1932.Google Scholar
Packard, A. (1972). Cephalopods and fish: the limits of convergence. Biological Reviews, 47, 241307.Google Scholar
Packard, A. (1974). Chromatophore fields in skin of Octopus. Journal of Physiology-London, 238, P38P40.Google Scholar
Packard, A. (1982). Morphogenesis of chromatophore patterns in cephalopods: are morphological and physiological ‘units’ the same? Malacologia, 23, 193201.Google Scholar
Packard, A. (1985). Sizes and distribution of chromatophores during post-embryonic development in cephalopods. Vie et Milieu, 35, 285298.Google Scholar
Packard, A. (1988a). Visual tactics and evolutionary strategies, in Cephalopods – Present and Past (ed. Wiedmann, J. & Kullmann, J.), pp. 89103. Stuttgart, Germany: Schweizerbart’sche Verlagsbuchhandlung.Google Scholar
Packard, A. (1988b). The skin of cephalopods (Coleoids): general and special adaptations, in The Mollusca, Vol. 11: Form and Function (ed. Trueman, E. R. & Clarke, M. R.), pp. 3767. San Diego: Academic Press.Google Scholar
Packard, A. (1995). Organization of cephalopod chromatophore systems: a neuromuscular image-generator, in Cephalopod Neurobiology (ed. Abbott, N. J., Williamson, R. & Maddock, L.), pp. 331368. Oxford: Oxford University Press.Google Scholar
Packard, A. & Brancato, D. (1993). Some responses of octopus chromatophores to light. Journal of Physiology London 459, P429P429.Google Scholar
Packard, A. & Hochberg, F. G. (1977). Skin patterning in Octopus and other genera. Symposia of the Zoological Society of London, 38, 191231.Google Scholar
Packard, A. & Sanders, G. (1969). What the octopus shows to the world. Endeavour, 28, 9299.Google Scholar
Packard, A. & Sanders, G. D. (1971). Body patterns of Octopus vulgaris and maturation of the response to disturbance. Animal Behaviour, 19, 780790.Google Scholar
Packard, A. & Wurtz, M. (1994). An Octopus, Ocythoe, with a swimbladder and triple jets. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 344, 261275.Google Scholar
Packard, A., Bone, Q. & Hignette, M. (1980). Breathing and swimming movements in a captive Nautilus. Journal of the Marine Biological Association of the United Kingdom, 60, 313327.Google Scholar
Packard, A., Karlsen, H. E. & Sand, O. (1990). Low frequency hearing in cephalopods. Journal of Comparative Physiology A, 166, 501505.Google Scholar
Palmer, B. W. & O’Dor, R. K. (1978). Changes in vertical migration patterns of captive Illex illecebrosus in varying light regimes and salinity gradients. Fisheries and Marine Service Technical Report, No. 833, 23.1–23.12.Google Scholar
Palmer, M. E., Calvé, M. R. & Adamo, S. A. (2006). Response of female cuttlefish Sepia officinalis (Cephalopoda) to mirrors and conspecifics: evidence for signaling in female cuttlefish. Animal Cognition, 9, 151155.Google Scholar
Papini, M. R. & Bitterman, M. E. (1991). Appetitive conditioning in Octopus cyanea. Journal of Comparative Psychology, 105, 107114.Google Scholar
Parker, G. A. (1970). Sperm competition and its evolutionary consequences in the insects. Biological Reviews, 45, 525567.Google Scholar
Parker, G. A. (1990). Sperm competition games: sneaks and extra-pair copulations. Proceedings of the Royal Society of London B, 242, 127133.Google Scholar
Parker, G. A. & Pizzari, T. (2010). Sperm competition and ejaculate economics. Biological Reviews, 85, 897934.Google Scholar
Parker, G. A., Simmons, L. W. & Kirk, H. (1990). Analyzing sperm competition data – simple models for predicting mechanisms. Behavioral Ecology and Sociobiology, 27, 5565.Google Scholar
Parry, M. (2000). A description of the nuchal organ, a possible photoreceptor, in Euprymna scolopes and other cephalopods. Journal of Zoology, 252, 163177.Google Scholar
Parry, M. (2006). Feeding behavior of two ommastrephid squids Ommastrephes bartramii and Sthenoteuthis oualaniensis off Hawaii. Marine Ecology Progress Series, 318, 229235.Google Scholar
Partridge, B. L. & Pitcher, T. (1980). The sensory basis of fish schools: relative roles of lateral line and vision. Journal of Comparative Physiology A, 135, 315325.Google Scholar
Pascual, E. (1978). Crecimiento y alimentación de tres generaciónes de Sepia officinalis en cultivo. Investigación Pesquera, 42, 421442.Google Scholar
Passarella, K. C. & Hopkins, T. L. (1991). Species composition and food habits of the micronektonic cephalopod assemblage in the Eastern Gulf of Mexico. Bulletin of Marine Science, 49, 638659.Google Scholar
Payne, N. L., Gillanders, B. M. & Semmens, J. (2011). Breeding durations as estimators of adult sex ratios and population size. Oecologia, 165, 341347.Google Scholar
Payne, N. L., Gillanders, B. M., Seymour, R. S. et al. (2010a). Accelerometry estimates field metabolic rate in giant Australian cuttlefish Sepia apama during breeding. Journal of Animal Ecology, 80, 422430.Google Scholar
Payne, N. L., Gillanders, B. M., Webber, D. M. & Semmens, J. M. (2010b). Interpreting diel activity patterns from acoustic telemetry: the need for controls. Marine Ecology Progress Series, 419, 295301.Google Scholar
Payne, R. J. H. (1998). Gradually escalating fights and displays: the cumulative assessment model. Animal Behaviour, 56, 651662.Google Scholar
Payne, R. J. H. & Pagel, M. (1996). Escalation and time costs in displays of endurance. Journal of Theoretical Biology, 183, 185193.Google Scholar
Payne, R. J. H. & Pagel, M. (1997). Why do animals repeat displays? Animal Behaviour, 54, 109119.Google Scholar
Pearce, J. M. (2008). Animal Learning and Cognition. Hove, UK: Psychology Press.Google Scholar
Pham, C. K., Carreira, G. P., Porteiro, F. M. et al. (2009). First description of spawning in a deep water loliginid squid, Loligo forbesi (Cephalopoda: Myopsida). Journal of the Marine Biological Association of the United Kingdom, 89, 171177.Google Scholar
Philips, M. & Austad, S. N. (1992). Animal communication and social evolution, in Interpretation and Explanation in the Study of Animal Behaviour, Vol. 1: Interpretation, Intentionality, and Communication (ed. Behoff, M. & Jamieson, D.), pp. 254268. Boulder, CO: Westview Press.Google Scholar
Pickford, G. E. & McConnaughey, B. H. (1949). The Octopus bimaculatus problem: a study in sibling species. Bulletin of the Bingham Oceanographic Collection, 12, 166.Google Scholar
Pieron, H. (1911). Contribution à la pyschologie du poulpe. Bulletin de l’Institut General Pyschologique, 11, 111119.Google Scholar
Pietrewicz, A. T. & Kamil, A. C. (1981). Search images and the detection of cryptic prey: an operant approach, in Foraging Behavior: Ecological, Ethological, and Psychological Approaches (ed. Kamil, A. C. & Sargent, T. D.), pp. 311331. New York, NY: Garland STPM Press.Google Scholar
Pignatelli, V., Temple, S. E., Chiou, T. H. et al. (2011). Behavioural relevance of polarization sensitivity as a target detection mechanism in cephalopods and fishes. Philosophical Transactions of the Royal Society B – Biological Sciences, 366, 734741.Google Scholar
Pimm, S. L., Lawton, J. H. & Cohen, J. E. (1991). Food web patterns and their consequences. Nature, 350, 669674.Google Scholar
Pitcher, T. J. (1983). Heuristic definitions of fish shoaling behaviour. Animal Behaviour, 31, 611612.Google Scholar
Pitcher, T. J. (1986). Functions of shoaling behavior in teleosts, in The Behaviour of Teleost Fishes (ed. Pitcher, T. J.), pp. 294337. Croon Helm: London.Google Scholar
Pitcher, T. J. (1993). The Behaviour of Teleost Fishes. London: Chapman & Hall.Google Scholar
Pitcher, T. J. & Parrish, J. K. (1993). Functions of shoaling behaviour in teleosts, in Behaviour of Teleost Fishes (ed. Pitcher, T. J.), pp. 363440. London: Chapman & Hall.Google Scholar
Pitcher, T. J. & Partridge, B. L. (1979). Fish school density and volume. Marine Biology, 54, 383394.Google Scholar
Pitcher, T. J. & Wyche, C. J. (1983). Predator-avoidance behaviours of sand-eel schools: why schools seldom split, in Predators and Prey in Fishes (ed. Noakes, D. L. G., Linquist, D. G., Helfman, G. S. & Ward, J. A.), pp. 193204. The Hague, Netherlands: Dr. W. Junk Publ.Google Scholar
Pitcher, T. J., Partridge, B. L. & Wardle, C. S. (1976). A blind fish can school. Science, 194, 963965.Google Scholar
Pliny, (Plinius, G. S.) (AD 77; this edition 1963). Natural History. London: Heinemann.Google Scholar
Ploger, B. J. & Yasukawa, K. (2003). Exploring Animal Behavior in Laboratory and Field: An Hypothesis-Testing Approach to the Development, Causation, Function, and Evolution of Animal Behavior: Elsevier.Google Scholar
Poirier, R., Chichery, R. & Dickel, L. (2004). Effects of rearing conditions on sand digging efficiency in juvenile cuttlefish. Behavioural Processes, 67, 273279.Google Scholar
Poirier, R., Chichery, R. & Dickel, L. (2005). Early experience and postembryonic maturation of body patterns in cuttlefish (Sepia officinalis). Journal of Comparative Psychology, 119, 230237.Google Scholar
Polese, G., Bertapelle, C. & Di Cosmo, A. (2015). Role of olfaction in Octopus vulgaris reproduction. General and Comparative Endocrinology, 210, 5562.Google Scholar
Polimanti, O. (1910). Les cephalopodes ont-ils une memorie? Archives de Psychologie Geneve, 10, 8487.Google Scholar
Porteiro, F. M., Martins, H. R. & Hanlon, R. T. (1990). Some observations on the behavior of adult squids, Loligo forbesi, in captivity. Journal of the Marine Biological Association of the United Kingdom, 70, 459472.Google Scholar
Pörtner, H. O. (2002). Environmental and functional limits to muscular exercise and body size in marine invertebrate athletes. Comparative Biochemistry and Physiology A – Molecular and Integrative Physiology, 133, 303321.Google Scholar
Preuss, T. & Budelmann, B. U. (1995a). Proprioceptive hair-cells on the neck of the squid Lolliguncula brevis – a sense organ in cephalopods for the control of head-to-body position. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 349, 153178.Google Scholar
Preuss, T. & Budelmann, B. U. (1995b). A dorsal light reflex in a squid. Journal of Experimental Biology, 198, 11571159.Google Scholar
Preuss, T. & Gilly, W. F. (2000). Role of prey-capture experience in the development of the escape response in the squid Loligo opalescens: a physiological correlate in an identified neuron. Journal of Experimental Biology, 203, 559565.Google Scholar
Pronk, R., Wilson, D. R. & Harcourt, R. (2010). Video playback demonstrates episodic personality in the gloomy octopus. Journal of Experimental Biology, 213, 10351041.Google Scholar
Prota, G. J., Ortonne, P., Voulot, C. et al. (1981). Occurrence and properties of tyrosinase in the ejected ink of cephalopods. Comparative Biochemistry and Physiology, 15, 453466.Google Scholar
Pumphrey, R. J. & Young, J. Z. (1938). The rates of conduction of nerve fibres of various diameters in cephalopods. Journal of Experimental Biology, 15, 453466.Google Scholar
Quetglas, A., Alemany, F., Carbonell, A., Merella, P. & Sanchez, P. (1999). Diet of the European flying squid Todarodes sagittatus (Cephalopoda: Ommastrephidae) in the Balearic Sea (western Mediterranean). Journal of the Marine Biological Association of the United Kingdom, 79, 479486.Google Scholar
Quinteiro, J., Baibai, T., Oukhattar, L. et al. (2011). Multiple paternity in the common octopus Octopus vulgaris (Cuvier, 1797), as revealed by microsatellite DNA analysis. Molluscan Research, 31, 1520.Google Scholar
Racovitza, E. G. (1894). Sur l’accouplement de quelques Céphalopodes Sepiola rondeletti (Leach), Rossia macrosoma (d. Ch.) et Octopus vulgaris (Lam.) [About the mating behavior of the cephalopods …]. Comptes Rendus Hebdomadaires des Seances de l’Academie des Sciences, Paris, Series D, 118, 722724.Google Scholar
Radakov, D. V. (1973). Schooling in the Ecology of Fish, Israel Programme for Scientific Translations.Google Scholar
Ramachandran, V. S., Tyler, C. W., Gregory, R. L., et al. (1996). Rapid adaptive camouflage in tropical flounders. Nature, 379, 815818.Google Scholar
Randall, J. E. (1967). Food habits of reef fishes of West Indies. Studies in Tropical Oceanography, 5, 665847.Google Scholar
Reynolds, J. D. (1996). Animal breeding systems. Trends in Ecology and Evolution, 11, 6872.Google Scholar
Richard, A. (1971). Action qualitative de la lumiere dans le determinisme du cycle sexuel chez le Cephalopode Sepia officinalis L. Comptes Rendus Hebdomadaires des Seances de l’Academie des Sciences, Paris, Series D, 272, 106109.Google Scholar
Ricklefs, R. E. & Miller, G. L. (1999). Ecology. New York: W.H. Freeman & Co.Google Scholar
Rigby, P. R. & Sakurai, Y. (2005). Multidimensional tracking of giant Pacific octopuses in northern Japan reveals unexpected foraging behaviour. Marine Technology Society Journal, 39, 6467.Google Scholar
Roberts, M. J., Barange, M., Lipinski, M. R. & Prowse, M. R. (2002). Direct hydroacoustic observations of chokka squid Loligo vulgaris reynaudii spawning activity in deep water. South African Journal of Marine Science – Suid-Afrikaanse Tydskrif vir Seewetenskap, 24, 387393.Google Scholar
Roberts, M. J., Downey, N. J. & Sauer, W. H. (2012). The relative importance of shallow and deep shelf spawning habitats for the South African chokka squid (Loligo reynaudii). ICES Journal of Marine Science, 69, 563571.Google Scholar
Robertson, J. D. (1994). Cytochalasin-D blocks touch learning in Octopus vulgaris. Proceedings of the Royal Society of London Series B – Biological Sciences, 258, 6166.Google Scholar
Robertson, J. D. & Lee, P. (1990). An electron microscopic and behavioral study of tactile learning and memory in Octopus vulgaris. Progress in Cell Research, 1, 287306.Google Scholar
Robertson, J. D., Bonaventura, J. & Kohm, A. (1995). Nitric oxide synthase inhibition blocks octopus touch learning without producing sensory or motor dysfunction. Proceedings of the Royal Society B – Biological Sciences, 261, 167172.Google Scholar
Robertson, J. D., Bonaventura, J. & Kohm, A. P. (1994). Nitric oxide is required for tactile learning in Octopus vulgaris. Proceedings of the Royal Society of London B – Biological Sciences, 256, 269273.Google Scholar
Robins, C. R., Bailey, R. M., Bond, C. E. et al. (1991). Common and scientific names of fishes from the United States and Canada, American Fisheries Society Special Publication. Bethesda, MD: American Fisheries Society.Google Scholar
Robinson, M. H. (1969). Defenses against visually hunting predators, in Evolutionary Biology (ed. Dobzhansky, T., Hecht, M. K. & Steere, W. C.), pp. 225259. New York, NY: Appleton-Century-Crofts Publ.Google Scholar
Robison, B. H. & Young, R. E. (1981). Bioluminescence in pelagic octopods. Pacific Science, 35, 3944.Google Scholar
Robison, B. H., Reisenbichler, K. R., Hunt, J. C. & Haddock, S. H. D. (2003). Light production by the arm tips of the deep-sea cephalopod Vampyroteuthis infernalis. Biological Bulletin, 205, 102109.Google Scholar
Rocha, F. & Guerra, A. (1996). Signs of an extended and intermittent terminal spawning in the squids Loligo vulgaris Lamarck and Loligo forbesi Steenstrup (Cephalopoda: Loliginidae). Journal of Experimental Marine Biology and Ecology, 207, 177189.Google Scholar
Rocha, F., Gonzalez, A. F., Segonzac, M. & Guerra, A. (2002). Behavioural observations of the cephalopod Vulcanoctopus hydrothermalis. Cahiers de Biologie Marine, 43, 299302.Google Scholar
Rocha, F., Guerra, A. & Gonzalez, A. F. (2001). A review of reproductive strategies in cephalopods. Biological Reviews, 76, 291304.Google Scholar
Rocha, L., Ross, R. & Kopp, G. (2012). Opportunistic mimicry by a Jawfish. Coral Reefs, 31, 285285.Google Scholar
Rodaniche, A. F. (1984). Iteroparity in the lesser Pacific striped octopus Octopus chierchiae. Bulletin of Marine Science, 35, 99104.Google Scholar
Rodaniche, A. F. (1991). Notes on the behavior of the larger Pacific striped octopus, an undescribed species of the genus Octopus. Bulletin of Marine Science, 49, 667.Google Scholar
Rodhouse, P. G., Prince, P. A., Clarke, M. R. & Murray, A. W. A. (1990). Cephalopod prey of the grey-headed albatross Diomedea chrysostoma. Marine Biology, 104, 353362.Google Scholar
Rodhouse, P. G., Swinfen, R. C. & Murray, A. W. A. (1988). Life cycle, demography and reproductive investment in the myopsid squid Alloteuthis subulata. Marine Ecology Progress Series, 45, 245253.Google Scholar
Rodrigues, M., Garci, M. E., Troncoso, J. S. & Guerra, A. (2010). Burying behaviour in the bobtail squid Sepiola atlantica (Cephalopoda: Sepiolidae). Italian Journal of Zoology, 77, 247251.Google Scholar
Rodrigues, M., Garci, M. E., Troncoso, J. S. & Guerra, A. (2011). Spawning strategy in Atlantic bobtail squid Sepiola atlantica (Cephalopoda: Sepiolidae). Helgoland Marine Research, 65, 4349.Google Scholar
Roeleveld, M. A. & Lipinski, M. R. (1991). The giant squid Architeuthis in southern African waters. Journal of Zoology (London), 224, 431477.Google Scholar
Roffe, T. (1975). Spectral perception in Octopus: a behavioral study. Vision Research, 15, 353356.Google Scholar
Romagny, S., Darmaillacq, A. S., Guibé, M., Bellanger, C. & Dickel, L. (2012). Feel, smell and see in an egg: emergence of perception and learning in an immature invertebrate, the cuttlefish embryo. Journal of Experimental Biology, 215, 41254130.Google Scholar
Romanini, M. G. (1952). Osservazioni sulla ialuronidasi delle ghiandole salivari anteriori e posteriori degli Octopodi. Pubblicazioni della Stazione Zoologica di Napoli, 23, 251270.Google Scholar
Roper, C. F. E. & Boss, K. J. (1982). The giant squid. Scientific American, 246, 96105.Google Scholar
Roper, C. F. E. & Brundage, W. L. (1972). Cirrate octopods with associated deep-sea organisms: new biological data based on deep benthic photographs (Cephalopoda). Smithsonian Contributions to Zoology, 121, 146.Google Scholar
Roper, C. F. E. & Hochberg, F. G. (1988). Behavior and systematics of cephalopods from Lizard Island, Australia, based on color and body patterns. Malacologia, 29, 153193.Google Scholar
Roper, C. F. E. & Vecchione, M. (1993). A geographic and taxonomic review of Taningia danae Joubin, 1931(Cephalopoda: Octopoteuthidae), with new records and observations on bioluminescence, in Recent Advances in Fisheries Biology (ed. Okutani, T., O’Dor, R. K. & Kubodera, T.), pp. 441456. Tokyo: Tokai University Press.Google Scholar
Roper, C. F. E. & Vecchione, M. (1996). In situ observations on Brachioteuthis beanii Verrill: paired behavior, probably mating (Cephalopoda, Oegopsida). American Malacological Bulletin, 13, 5560.Google Scholar
Roper, C. F. E. & Young, R. E. (1975). Vertical distribution of pelagic cephalopods. Smithsonian Contributions to Zoology, 209, 51.Google Scholar
Roper, C. F. E., Sweeney, M. J. & Nauen, C. E. (1984). FAO Species Catalogue. Cephalopods of the World. An annotated and illustrated catalogue of species of interest to fisheries. FAO Fisheries Synopsis, 3, 1277.Google Scholar
Roper, C. F. E., Young, R. E. & Voss, G. L. (1969). An illustrated key to the families of the order Teuthoidea (Cephalapoda). Smithsonian Contributions to Zoology, 13, 132.Google Scholar
Rosa, R. & Seibel, B. A. (2010a). Slow pace of life of the Antarctic colossal squid. Journal of the Marine Biological Association of the United Kingdom, 90, 13751378.Google Scholar
Rosa, R. & Seibel, B. A. (2010b). Voyage of the argonauts in the pelagic realm: physiological and behavioural ecology of the rare Paper Nautilus, Argonauta nouryi. ICES Journal of Marine Science: Journal du Conseil, 67, 14941500.Google Scholar
Rosa, R., O’Dor, R. & Pierce, G. J. (2013). Advances in Squid Biology, Ecology and Fisheries. Part 1: Myopsid Squids. Hauppauge, NY: Nova Publishers.Google Scholar
Ross, D. M. (1971). Protection of hermit crabs (Dardanus spp.) from octopus by commensal sea anemones (Calliactis spp.). Nature, 230, 401402.Google Scholar
Ross, D. M. & Boletzky, S. von (1979). The association between the pagurid Dardanus arrosor and the actinian Calliactis parasitica. Recovery of activity in ‘inactive’ D. arrosor in the presence of cephalopods. Marine Behaviour and Physiology, 6, 175184.Google Scholar
Rowell, C. H. F. & Wells, M. J. (1961). Retinal orientation and the discrimination of polarized light by octopuses. Journal of Experimental Biology, 38, 827831.Google Scholar
Rowland, H. M. (2009). From Abbott Thayer to the present day: what have we learned about the function of countershading? Philosophical Transactions of the Royal Society B – Biological Sciences, 364, 519527.Google Scholar
Royan, A., Muir, A. P. & Downie, J. R. (2010). Variability in escape trajectory in the Trinidadian stream frog and two treefrogs at different life-history stages. Canadian Journal of Zoology – Revue Canadienne de Zoologie, 88, 922934.Google Scholar
Ruth, P., Schmidtberg, H., Westermann, B. & Schipp, R. (2002). The sensory epithelium of the tentacles and the rhinophore of Nautilus pompilius L. (Cephalopoda, Nautiloidea). Journal of Morphology, 251, 239255.Google Scholar
Ruxton, G. D., Sherratt, T. N. & Speed, M. P. (2004). Avoiding Attack: The Evolutionary Ecology of Crypsis, Warning Signals, and Mimicry. Oxford: Oxford University Press.Google Scholar
Saibil, H. R. (1990). Cell and molecular biology of photoreceptors. Seminars in the Neurosciences, 2, 1523.Google Scholar
Saibil, H. R. & Hewat, E. (1987). Ordered transmembrane and extracellular structure in squid photoreceptor microvilli. Journal of Cell Biology, 105, 1928.Google Scholar
Saidel, W. M., Lettvin, J. Y. & MacNichol, E. F. J. (1983). Processing of polarized light by squid photoreceptors. Nature, 304, 534536.Google Scholar
Saidel, W. M., Shashar, N., Schmolesky, M. T. & Hanlon, R. T. (2005). Discriminative responses of squid (Loligo pealeii) photoreceptors to polarized light. Comparative Biochemistry and Physiology A, 142, 340346.Google Scholar
Sakamoto, W., Naito, Y., Huntley, A. C. & Le Boeuf, B. J. (1989). Daily gross energy requirements of a female northern elephant seal Mirounga angustirostris at sea. Nippon Suisan Gakkaishi, 55, 20572063.Google Scholar
Sakurai, Y., Bower, J. R. & Ikeda, Y. (2003). Reproductive characteristics of the ommastrephid squid Todarodes pacificus, in Modern Approaches to Assess Maturity and Fecundity of Warm- and Cold-Water Fish and Squids, Vol. 12 (ed. Kjesbu, O. S., Hunter, J. R. & Witthames, P. R.), pp. 105115. Bergen, Norway: Institute of Marine Research.Google Scholar
Sakurai, Y., Kidokoro, H., Yamashita, N. et al. (2013). Todarodes pacificus, Japanese common squid, in Advances in Squid Biology, Ecology and Fisheries. Part II (ed. Rosa, R., Pierce, G. J. & O’Dor, R.), pp. 249–271.Google Scholar
Samson, J. E., Mooney, T. A., Gussekloo, S. W. S. & Hanlon, R. T. (2014). Graded behavioral responses and habituation to sound in the common cuttlefish Sepia officinalis. Journal of Experimental Biology, 217, 43474355.Google Scholar
Sanchez, P. (2003). Cephalopods from off the Pacific coast of Mexico: biological aspects of the most abundant species. Scientia Marina, 67, 8190.Google Scholar
Sanders, F. K. & Young, J. Z. (1940). Learning and other functions of the higher nervous centres of Sepia. Journal of Neurophysiology, 3, 501526.Google Scholar
Sanders, G. D. (1970). Long-term memory of a tactile discrimination in Octopus vulgaris and the effect of vertical lobe removal. Brain Research, 20, 5973.Google Scholar
Sanders, G. D. (1975). The Cephalopods, in Invertebrate Learning, Vol. 3: Cephalopods and Echinoderms (ed. Corning, W. C., Dyal, J. A. & Willows, A. O. D.), pp. 1101. New York, NY: Plenum Press.Google Scholar
Sato, N., Kasugai, T., Ikeda, Y. & Munehara, H. (2010). Structure of the seminal receptacle and sperm storage in the Japanese pygmy squid. Journal of Zoology, 282, 151156.Google Scholar
Sato, N., Kasugai, T. & Munehara, H. (2013). Sperm transfer or spermatangia removal: postcopulatory behaviour of picking up spermatangium by female Japanese pygmy squid. Marine Biology, 160, 553561.Google Scholar
Sato, N., Kasugai, T. & Munehara, H. (2014). Female pygmy squid cryptically favour small males and fast copulation as observed by removal of spermatangia. Evolutionary Biology, 41, 221228.Google Scholar
Sato, N., Yoshida, M.-A., Fujiwara, E. & Kasugai, T. (2013). High-speed camera observations of copulatory behaviour in Idiosepius paradoxus: function of the dimorphic hectocotyli. Journal of Molluscan Studies, 79, 183186.Google Scholar
Sauer, W. H. H. (1995). The impact of fishing on chokka squid Loligo vulgaris reynaudii concentrations on inshore spawning grounds in the South-Eastern Cape, South Africa. South African Journal of Marine Science – Suid-Afrikaanse Tydskrif vir Seewetenskap, 16, 185193.Google Scholar
Sauer, W. H. H. & Smale, M. J. (1993). Spawning behaviour of Loligo vulgaris reynaudii in shallow coastal waters of the South Eastern Cape, South Africa, in Recent Advances in Fisheries Biology (ed. Okutani, T., O’Dor, R. K. & Kubodera, T.), pp. 489498. Tokyo: Tokai University Press.Google Scholar
Sauer, W. H. H., Roberts, M. J., Lipinski, M. R. et al. (1997). Choreography of the squid’s ‘nuptial dance’. Biological Bulletin, 192, 203207.Google Scholar
Sauer, W. H. H., Smale, M. J. & Lipinski, M. R. (1992). The location of spawning grounds, spawning and schooling behaviour of the squid Loligo vulgaris reynaudii (Cephalopoda: Myopsida) off the Eastern Cape Coast, South Africa. Marine Biology, 114, 97107.Google Scholar
Saunders, W. B. (1983). Natural rates of growth and longevity of Nautilus belauensis. Paleobiology, 9, 280288.Google Scholar
Saunders, W. B. (1984). The role and status of Nautilus in its natural habitat: evidence from deep-water remote camera photosequences. Paleobiology, 10, 469486.Google Scholar
Saunders, W. B. (1985). Studies of living Nautilus in Palau. National Geographic Society Research Reports, 18, 669682.Google Scholar
Saunders, W. B. & Landman, N. H. (2009). Nautilus: The Biology and Paleobiology of a Living Fossil. New York: Plenum Press.Google Scholar
Saunders, W. B. & Spinosa, C. (1979). Nautilus movement and distribution in Palau, Western Caroline Islands. Science, 204, 11991201.Google Scholar
Saunders, W. B. & Ward, P. D. (1987). Ecology, distribution and population characteristics of Nautilus, in Nautilus: The Biology and Paleobiology of a Living Fossil (ed. Saunders, W. B. & Landman, N. H.), pp. 137162. New York: Plenum Press.Google Scholar
Saunders, W. B., Knight, R. L. & Bond, P. N. (1991). Octopus predation on Nautilus: evidence from Papua New Guinea. Bulletin of Marine Science, 49, 280287.Google Scholar
Saunders, W. B., Spinosa, C. & Davis, L. E. (1987). Predation on Nautilus, in Nautilus: The Biology and Paleobiology of a Living Fossil (ed. Saunders, W. B. & Landman, N. H.), pp. 201212. New York: Plenum Press.Google Scholar
Schäfer, W. (1936). Bau, Entwicklung und Farbenentstehung bei den Flitterzellen von Sepia officinalis. Zeitschrift für Zellforschung und Mikroskopische Anatomie, 27, 222245.Google Scholar
Schäfer, W. (1956). Die Schutzwirkung der Tintenfisch-Tinte. Natur und Volk, 86, 2426.Google Scholar
Scheel, D. & Anderson, R. (2012). Variability in the diet specialization of Enteroctopus dofleini (Cephalopoda: Octopodidae) in the eastern Pacific examined from midden contents. American Malacological Bulletin, 30, 267279.Google Scholar
Scheel, D. & Bisson, L. (2012). Movement patterns of giant Pacific octopuses, Enteroctopus dofleini (Walker, 1910). Journal of Experimental Marine Biology and Ecology, 416–417, 2131.Google Scholar
Schiller, P. H. (1949). Delayed detour response in the octopus. Journal of Comparative and Physiological Psychology, 42, 220225.Google Scholar
Schmitt, R. J. (1982). Consequences of dissimilar defenses against predation in a subtidal marine community. Ecology, 63, 15881601.Google Scholar
Schmitt, R. J. (1987). Indirect interactions between prey: apparent competition, predator aggregation, and habitat segregation. Ecology, 68, 18871897.Google Scholar
Schnell, A. K. (2014). Signalling, mating and conflict resolution in the Giant Australian cuttlefish, Sepia apama. Ph.D. Thesis. Sydney, Australia: Macquarie University.Google Scholar
Schnell, A. K., Smith, C. L., Hanlon, R. T. & Harcourt, R. (2015). Giant Australian cuttlefish use mutual assessment to resolve male–male contests. Animal Behaviour, 107, 3140.Google Scholar
Schoener, T. W. (1974). Resource partitioning in ecological communities. Science, 185, 2739.Google Scholar
Scott, W. B. & Tibbo, S. N. (1968). Food and feeding habits of swordfish, Xiphias gladius, in the western North Atlantic. Journal of the Fisheries Research Board of Canada, 25, 903919.Google Scholar
Segawa, S. (1987). Life history of the oval squid, Sepioteuthis lessoniana, in Kominato and adjacent waters central Honshu, Japan. Journal of the Tokyo University of Fisheries, 74, 67105.Google Scholar
Segawa, S., Izuka, T., Tamashiro, T. & Okutani, T. (1993). A note on mating and egg deposition by Sepioteuthis lessoniana in Ishigaki Island, Okinawa, Southwestern Japan. Venus: Japanese Journal of Malacology, 52, 101108.Google Scholar
Segawa, S., Yang, W. T., Marthy, H. J. & Hanlon, R. T. (1988). Illustrated embryonic stages of the eastern Atlantic squid Loligo forbesi. Veliger, 30, 230243.Google Scholar
Seibel, B. A. (2011). Critical oxygen levels and metabolic suppression in oceanic oxygen minimum zones. Journal of Experimental Biology, 214, 326336.Google Scholar
Seibel, B. A., Hochberg, F. G. & Carlini, D. B. (2000). Life history of Gonatus onyx (Cephalopoda: Teuthoidea): deep-sea spawning and post-spawning egg care. Marine Biology, 137, 519526.Google Scholar
Seibel, B. A., Robison, B. H. & Haddock, S. H. D. (2005). Post-spawning egg care by a squid. Nature, 438, 929.Google Scholar
Seidou, M., Sugahara, M., Uchiyama, H. et al. (1990). On the three visual pigments in the retina of the firefly squid, Watasenia scintillans. Journal of Comparative Physiology A, 166, 769773.Google Scholar
Sereni, E. & Young, J. Z. (1932). Nervous degeneration and regeneration in cephalopods. Pubblicazioni della Stazione Zoologica di Napoli, 12, 173208.Google Scholar
Seyfarth, R. M. & Cheney, D. L. (2003). Signalers and receivers in animal communication. Annual Review of Psychology, 54, 145173.Google Scholar
Shashar, N. & Cronin, T. W. (1996). Polarization contrast vision in octopus. Journal of Experimental Biology, 199, 9991004.Google Scholar
Shashar, N. & Hanlon, R. T. (1997). Squids (Loligo pealei and Euprymna scolopes) can exhibit polarized light patterns produced by their skin. Biological Bulletin, 193, 207208.Google Scholar
Shashar, N. & Hanlon, R. T. (2013). Spawning behavior dynamics at communal egg beds in the squid Doryteuthis (Loligo) pealeii. Journal of Experimental Marine Biology and Ecology, 447, 6574.Google Scholar
Shashar, N., Hagan, R., Boal, J. G. & Hanlon, R. T. (2000). Cuttlefish use polarization sensitivity in predation on silvery fish. Vision Research, 40, 7175.Google Scholar
Shashar, N., Hanlon, R. T. & Petz, A. D. (1998). Polarization vision helps detect transparent prey. Nature, 393, 222223.Google Scholar
Shashar, N., Harosi, F. I., Banaszak, A. T. & Hanlon, R. T. (1998). UV radiation blocking compounds in the eye of the cuttlefish Sepia officinalis. Biological Bulletin, 195, 187188.Google Scholar
Shashar, N., Johnsen, S., Lerner, A. et al. (2011). Underwater linear polarization: physical limitations to biological functions. Philosophical Transactions of the Royal Society B – Biological Sciences, 366, 649654.Google Scholar
Shashar, N., Milbury, C. A. & Hanlon, R. T. (2002). Polarization vision in cephalopods: neuroanatomical and behavioral features that illustrate aspects of form and function. Marine and Freshwater Behaviour and Physiology, 35, 5768.Google Scholar
Shashar, N., Rutledge, P. S. & Cronin, T. W. (1996). Polarization vision in cuttlefish – a concealed communication channel? Journal of Experimental Biology, 199, 20772084.Google Scholar
Shaw, E. (1978). Schooling fishes. American Scientist, 66, 166175.Google Scholar
Shaw, P. W. (1997). Polymorphic microsatellite markers in a cephalopod: the veined squid Loligo forbesi. Molecular Ecology, 6, 297298.Google Scholar
Shaw, P. W. (2002). Past, present and future applications of DNA-based markers in cephalopod biology: workshop report. Bulletin of Marine Science, 71, 6778.Google Scholar
Shaw, P. W. (2003). Polymorphic microsatellite DNA markers for the assessment of genetic diversity and paternity testing in the giant cuttlefish, Sepia apama (Cephalopoda). Conservation Genetics, 4, 533535.Google Scholar
Shaw, P. W. & Boyle, P. R. (1997). Multiple paternity within the brood of single females of Loligo forbesi (Cephalopoda: Loliginidae), demonstrated with microsatellite DNA markers. Marine Ecology Progress Series, 160, 279282.Google Scholar
Shaw, P. W. & Sauer, W. H. H. (2004). Multiple paternity and complex fertilisation dynamics in the squid Loligo vulgaris reynaudii. Marine Ecology Progress Series, 270, 173179.Google Scholar
Shaw, P. W., Arkhipkin, A. I., Adcock, G. J. et al. (2004). DNA markers indicate that distinct spawning cohorts and aggregations of Patagonian squid, Loligo gahi, do not represent genetically discrete subpopulations. Marine Biology, 144, 961970.Google Scholar
Shaw, P. W., Hendrickson, L., McKeown, N. J. et al. (2010). Discrete spawning aggregations of loliginid squid do not represent genetically distinct populations. Marine Ecology Progress Series, 408, 117127.Google Scholar
Shaw, P. W., Hendrickson, L., McKeown, N. J. et al. (2012). Population structure of the squid Doryteuthis (Loligo) pealeii on the eastern coast of the USA: reply to Gerlach et al. (2012). Marine Ecology Progress Series, 450, 285287.Google Scholar
Shchetinnikov, A. S. (1992). Feeding spectrum of squid Sthenoteuthis oualaniensis (Oegopsida) in the Eastern Pacific. Journal of the Marine Biological Association of the United Kingdom, 72, 849860.Google Scholar
Shears, J. (1988). The use of a sand-coat in relation to feeding and diel activity in the sepiolid squid Euprymna scolopes. Malacologia, 29, 121133.Google Scholar
Shettleworth, S. J. (2010). Cognition, Evolution and Behaviour. Oxford: Oxford University Press.Google Scholar
Sheumack, D. D., Howden, M. E. H., Spence, I. & Quinn, R. J. (1978). Maculotoxin: a neurotoxin for the venom glands of the octopus Hapalochlaena maculosa identified as tetrodotoxin. Science, 199, 188189.Google Scholar
Shigeno, S., Sasaki, T., Moritaki, T. et al. (2008). Evolution of the cephalopod head complex by assembly of multiple molluscan body parts: evidence from Nautilus embryonic development. Journal of Morphology, 269, 117.Google Scholar
Shohet, A., Baddeley, O., Anderson, J. & Osorio, D. (2007). Cuttlefish camouflage: a quantitative study of patterning. Biological Journal of the Linnean Society, 92, 335345.Google Scholar
Shohet, A. J., Baddeley, R. J., Anderson, J. C., Kelman, E. J. & Osorio, D. (2006). Cuttlefish response to visual orientation of substrates, water flow and a model of motion camouflage. Journal of Experimental Biology, 209, 47174723.Google Scholar
Shomrat, T., Graindorge, N., Bellanger, C., Fiorito, G., Loewenstein, Y. & Hochner, B. (2011). Alternative sites of synaptic plasticity in two homologous ‘fan-out fan-in’ learning and memory networks. Current Biology, 21, 17731782.Google Scholar
Shomrat, T., Zarrella, I., Fiorito, G. & Hochner, B. (2008). The octopus vertical lobe modulates short-term learning rate and uses LTP to acquire long-term memory. Current Biology, 18, 337342.Google Scholar
Shuster, S. M. & Wade, M. J. (2003). Mating Systems and Strategies. Princeton, NJ: Princeton University Press.Google Scholar
Sifner, S. K. & Vrgoc, N. (2004). Population structure, maturation and reproduction of the European squid, Loligo vulgaris, in the Central Adriatic Sea. Fisheries Research, 69, 239249.Google Scholar
Simmons, L. W. (2005). The evolution of polyandry: sperm competition, sperm selection, and offspring viability. Annual Review of Ecology Evolution and Systematics, 36, 125146.Google Scholar
Singley, C. T. (1982). Histochemistry and fine structure of the ectodermal epithelium of the sepiolid squid Euprymna scolopes. Malacologia, 23, 177192.Google Scholar
Singley, C. T. (1983). Euprymna scolopes, in Cephalopod Life Cycles, Vol. 1 (ed. Boyle, P. R.), pp. 6974. London: Academic Press.Google Scholar
Sinn, D. L. & Moltschaniwskyj, N. A. (2005). Personality traits in dumpling squid (Euprymna tasmanica): context-specific traits and their correlation with biological characteristics. Journal of Comparative Psychology, 119, 99110.Google Scholar
Sinn, D. L., Apiolaza, L. A. & Moltschaniwskyj, N. A. (2006). Heritability and fitness-related consequences of squid personality traits. Journal of Evolutionary Biology, 19, 14371447.Google Scholar
Sinn, D. L., Gosling, S. D. & Moltschaniwskyj, N. A. (2008). Development of shy/bold behaviour in squid: context-specific phenotypes associated with developmental plasticity. Animal Behaviour, 75, 433442.Google Scholar
Sinn, D. L., Moltschaniwskyj, N., Wapstra, E. & Dall, S. (2010). Are behavioral syndromes invariant? Spatiotemporal variation in shy/bold behavior in squid. Behavioral Ecology and Sociobiology, 64, 693702.Google Scholar
Sinn, D. L., Perrin, N. A., Mather, J. A. & Anderson, R. C. (2001). Early temperamental traits in an octopus (Octopus bimaculoides). Journal of Comparative Psychology, 115, 351364.Google Scholar
Skelhorn, J., Rowland, H. M. & Ruxton, G. D. (2010). The evolution and ecology of masquerade. Biological Journal of the Linnean Society, 99, 18.Google Scholar
Skelhorn, J., Rowland, H. M., Speed, M. P. & Ruxton, G. D. (2010). Masquerade: camouflage without crypsis. Science, 327, 51.Google Scholar
Slater, P. J. B. (1983). The study of communication, in Communication (ed. Halliday, T. R. & Slater, P. J. B.). Oxford: Blackwell Scientific Publications.Google Scholar
Smale, M. J. (1996). Cephalopods as prey. 4. Fishes. Philosophical Transactions of the Royal Society of London Series B – Biological Sciences, 351, 10671081.Google Scholar
Smale, M. J., Sauer, W. H. H. & Hanlon, R. T. (1995). Attempted ambush predation on spawning squids Loligo vulgaris reynaudii by benthic pyjama sharks, Poroderma africanum, off South Africa. Journal of the Marine Biological Association of the United Kingdom, 75, 739742.Google Scholar
Smith, R. L. (1984). Sperm Competition and the Evolution of Animal Mating Systems. Orlando: Academic Press.Google Scholar
Smith, W. J. (1997). The behavior of communicating, after twenty years, in Communication (ed. Owings, D. H., Beecher, M. D. & Thompson, N. S.). New York: Plenum Press.Google Scholar
Sneddon, L. U. (2009). Pain perception in fish: indicators and endpoints. Ilar Journal, 50, 338342.Google Scholar
Soeda, J. (1965). Migration of the ‘Surume’ squid Ommastrephes sloani pacificus (Steenstrup), in the coastal waters of Japan. Fisheries Research Board of Canada, 533, 138.Google Scholar
Soucier, C. P. & Basil, J. A. (2008). Chambered nautilus (Nautilus pompilius pompilius) responds to underwater vibrations. Americal Malacological Bulletin, 24, 311.Google Scholar
Squires, H. J. (1957). Squid, Illex illecebrosus (Lesueur), in the Newfoundland fishing area. Journal of the Fisheries Research Board of Canada, 14, 693728.Google Scholar
Squires, Z. E., Norman, M. D. & Stuart-Fox, D. (2013). Mating behaviour and general spawning patterns of the southern dumpling squid Euprymna tasmanica (Sepiolidae): a laboratory study. Journal of Molluscan Studies, 79, 263269.Google Scholar
Squires, Z. E., Wong, B. B. M., Norman, M. D. & Stuart-Fox, D. (2012). Multiple fitness benefits of polyandry in a cephalopod. PLoS One, 7, e37074.Google Scholar
Staaf, D. J., Camarillo-Coop, S., Haddock, S. H. D. et al. (2008). Natural egg mass deposition by the Humboldt squid (Dosidicus gigas) in the Gulf of California and characteristics of hatchlings and paralarvae. Journal of the Marine Biological Association of the United Kingdom, 88, 759770.Google Scholar
Staples, J. F., Webber, D. M. & Boutilier, R. G. (2003). Environmental hypoxia does not constrain the diurnal depth distribution of free-swimming Nautilus pompilius. Physiological and Biochemical Zoology, 76, 644651.Google Scholar
Stark, K. E., Jackson, G. D. & Lyle, J. M. (2005). Tracking arrow squid movements with an automated acoustic telemetry system. Marine Ecology Progress Series, 299, 167177.Google Scholar
Staudinger, M. D. (2006). Seasonal and size-based predation on two species of squid by four fish predators on the Northwest Atlantic continental shelf. Fishery Bulletin, 104, 605615.Google Scholar
Staudinger, M. D. & Juanes, F. (2010a). Feeding tactics of a behaviorally plastic predator, summer flounder (Paralichthys dentatus). Journal of Sea Research, 64, 6875.Google Scholar
Staudinger, M. D. & Juanes, F. (2010b). Size-dependent susceptibility of longfin inshore squid (Loligo pealeii) to attack and capture by two predators. Journal of Experimental Marine Biology and Ecology, 393, 106113.Google Scholar
Staudinger, M. D., Buresch, K. C., Mathger, L. M. et al. (2013a). Defensive responses of cuttlefish to different teleost predators. Biological Bulletin, 225, 161174.Google Scholar
Staudinger, M. D., Hanlon, R. T. & Juanes, F. (2011). Primary and secondary defenses of squid to cruising and ambush fish predators: variable tactics and their survival value. Animal Behaviour, 81, 585594.Google Scholar
Staudinger, M. D., Juanes, F., Salmon, B. & Teffer, A. K. (2013b). The distribution, diversity, and importance of cephalopods in top predator diets from offshore habitats of the Northwest Atlantic Ocean. Deep-Sea Research Part II –Topical Studies in Oceanography, 95, 182192.Google Scholar
Stearns, S. C. (1992). The Evolution of Life Histories. Oxford: Oxford University Press.Google Scholar
Stella, M. P. & Kier, W. M. (2004). The morphology and mechanics of octopus arms: inspiration for novel robotics. Integrative and Comparative Biology, 44, 645645.Google Scholar
Stephens, P. R. & Young, J. Z. (1982). The statocyst of the squid Loligo. Journal of Zoology (London), 197, 241266.Google Scholar
Stevens, M. (2005). The role of eyespots as anti-predator mechanisms, principally demonstrated in the Lepidoptera. Biological Reviews, 80, 573588.Google Scholar
Stevens, M. (2013). Sensory Ecology, Behaviour and Evolution. Oxford: Oxford University Press.Google Scholar
Stevens, M. & Cuthill, I. C. (2006). Disruptive coloration, crypsis and edge detection in early visual processing. Proceedings of the Royal Society B – Biological Sciences 273, 21412147.Google Scholar
Stevens, M. & Merilaita, S. (2009a). Animal camouflage: current issues and new perspectives. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 364, 423557.Google Scholar
Stevens, M. & Merilaita, S. (2009b). Defining disruptive coloration and distinguishing its functions. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 364, 481488.Google Scholar
Stevens, M. & Merilaita, S. (ed.) (2011a). Animal Camouflage: Mechanisms and Function. Cambridge: Cambridge University Press.Google Scholar
Stevens, M. & Merilaita, S. (2011b). Animal camouflage: function and mechanisms, in Animal Camouflage: Mechanisms and Function (ed. Stevens, M. & Merilaita, S.), pp. 116. Cambridge: Cambridge University Press.Google Scholar
Stevens, M. & Ruxton, G. D. (2014). Do animal eyespots really mimic eyes? Current Zoology, 60, 2636.Google Scholar
Stevens, M., Cuthill, I. C., Windsor, A. M. M. & Walker, H. J. (2006). Disruptive contrast in animal camouflage. Proceedings of the Royal Society B: Biological Sciences, 273, 24332438.Google Scholar
Stevens, M., Graham, J., Winney, I. S. & Cantor, A. (2008). Testing Thayer’s hypothesis: can camouflage work by distraction? Biology Letters, 4, 648650.Google Scholar
Stevens, M., Marshall, K. L. A., Troscianko, J. et al. (2013). Revealed by conspicuousness: distractive markings reduce camouflage. Behavioral Ecology, 24, 213222.Google Scholar
Stevens, M., Stubbins, C. L. & Hardman, C. J. (2008). The anti-predator function of ‘eyespots’ on camouflaged and conspicuous prey. Behavioral Ecology and Sociobiology, 62, 17871793.Google Scholar
Stevens, M., Winney, I. S., Cantor, A. & Graham, J. (2009). Outline and surface disruption in animal camouflage. Proceedings of the Royal Society B – Biological Sciences, 276, 781786.Google Scholar
Stevenson, J. A. (1934). On the behaviour of the long-finned squid (Loligo pealii (Lesueur)). Canadian Field Naturalist, 48, 47.Google Scholar
Stewart, J. S., Field, J. C., Markaida, U. & Gilly, W. F. (2013). Behavioral ecology of jumbo squid (Dosidicus gigas) in relation to oxygen minimum zones. Deep-Sea Research Part II – Topical Studies in Oceanography, 95, 197208.Google Scholar
Stewart, J. S., Hazen, E. L., Foley, D. G., Bograd, S. J. & Gilly, W. F. (2012). Marine predator migration during range expansion: Humboldt squid Dosidicus gigas in the northern California Current System. Marine Ecology Progress Series, 471, 135150.Google Scholar
Stillwell, C. E. & Kohler, N. E. (1985). Food and feeding ecology of the swordfish Xiphias gladius in the western north-Atlantic Ocean with estimates of daily ration. Marine Ecology Progress Series, 22, 239247.Google Scholar
Strand, S. (1988). Following behavior: interspecific foraging associations among Gulf of California reef fishes. Copeia, 1988, 351357.CrossRefGoogle Scholar
Strugnell, J., Jackson, J., Drummond, A. J. & Cooper, A. (2006). Divergence time estimates for major cephalopod groups: evidence from multiple genes. Cladistics, 22, 8996.Google Scholar
Strugnell, J., Norman, M., Jackson, J., Drummond, A. J. & Cooper, A. (2005). Molecular phylogeny of coleoid cephalopods (Mollusca: Cephalopoda) using a multigene approach; the effect of data partitioning on resolving phylogenies in a Bayesian framework. Molecular Phylogenetics and Evolution, 37, 426441.Google Scholar
Stürmer, W. (1985). A small coleoid cephalopod with soft parts from the Lower Devonian discovered using radiography. Nature, 318, 5355.CrossRefGoogle Scholar
Suboski, M. D., Muir, D. & Hall, D. (1993). Social learning in invertebrates. Science, 259, 16281629.Google Scholar
Sugawara, K., Katagiri, Y. & Tomita, T. (1971). Polarized light respones from octopus single reticula cells. Journal of the Faculty of Sciences, Hokkaido University, 17, 581586.Google Scholar
Sugimoto, C., Yanagisawa, R., Nakajima, R. & Ikeda, Y. (2013). Observations of schooling behaviour in the oval squid Sepioteuthis lessoniana in coastal waters of Okinawa Island. Marine Biodiversity Records, 6, e34.Google Scholar
Sumbre, G., Fiorito, G., Flash, T. & Hochner, B. (2005). Neurobiology: motor control of flexible octopus arms. Nature, 433, 595596.Google Scholar
Sumbre, G., Fiorito, G., Flash, T. & Hochner, B. (2006). Octopuses use a human-like strategy to control precise point-to-point arm movements. Current Biology, 16, 767772.Google Scholar
Sumbre, G., Gutfreund, Y., Fiorito, G., Flash, T. & Hochner, B. (2001). Control of octopus arm extension by a peripheral motor program. Science, 293, 18451848.CrossRefGoogle ScholarPubMed
Summers, W. C. (1983). Loligo pealei, in Cephalopod Life Cycles, Vol. I: Species Accounts (ed. Boyle, P. R.), New York: Academic Press, pp. 115142.Google Scholar
Summers, W. C. (1985). Comparative life history adaptations of some myopsid and sepiolid squids. NAFO Scientific Council Studies, 9, 139142.Google Scholar
Sundermann, G. (1983). The fine structure of epidermal lines on arms and head of postembryonic Sepia officinalis and Loligo vulgaris (Mollusca, Cephalopoda). Cell and Tissue Research, 232, 669677.Google Scholar
Sundermann, G. (1990). Development and hatching state of ectodermal vesicle organs in the head of Sepia officinalis, Loligo vulgaris and Loligo forbesi (Cephalopoda, Decabrachia). Zoomorphology, 109, 343352.Google Scholar
Sutherland, N. S. (1957a). Visual discrimination of orientation by Octopus. British Journal of Psychology, 48, 5571.Google Scholar
Sutherland, N. S. (1957b). Visual discrimination of orientation and shape by the octopus. Nature, 179, 1113.CrossRefGoogle Scholar
Sutherland, N. S. (1958). Visual discrimination of shape by Octopus. Squares and triangles. Quarterly Journal of Experimental Psychology, 10, 4047.Google Scholar
Sutherland, N. S. (1959). A test of a theory of shape discrimination in Octopus vulgaris Lamarck. Journal of Comparative and Physiological Psychology, 52, 135141.Google Scholar
Sutherland, N. S. (1960). Theories of shape discrimination in Octopus. Nature, 186, 840844.CrossRefGoogle ScholarPubMed
Sutherland, N. S. (1962). Visual discrimination of shape by Octopus: squares and crosses. Journal of Comparative and Physiological Psychology, 55, 939943.Google Scholar
Sutherland, N. S. (1963). Shape discrimination and receptive fields. Nature, 197, 118122.Google Scholar
Sutherland, N. S. (1968). Outlines of a theory of visual pattern recognition in animals and man. Proceedings of the Royal Society of London B, 171, 297317.Google Scholar
Sutherland, N. S. (1969). Shape discrimination in rat, octopus and goldfish: a comparative study. Journal of Comparative and Physiological Psychology, 67, 160176.CrossRefGoogle ScholarPubMed
Sutherland, N. S. & Carr, A. E. (1963). The visual discrimination of shape by Octopus: the effects of stimulus size. Quarterly Journal of Experimental Psychology, 15, 225235.Google Scholar
Sutherland, N. S. & Mackintosh, N. J. (1971). Mechanisms of Animal Discrimination Learning. New York: Academic Press.Google Scholar
Sutherland, N. S. & Muntz, W. R. A. (1959). Simultaneous discrimination training and preferred directions of motion in visual discrimination of shape in Octopus vulgaris Lamarck. Pubblicazioni della Stazione Zoologica di Napoli, 31, 109126.Google Scholar
Sutherland, N. S., Mackintosh, N. J. & Mackintosh, J. (1963). Simultaneous discrimination training in Octopus and transfer of a discrimination along a continuum. Journal of Comparative and Physiological Psychology, 56, 150156.Google Scholar
Sutherland, N. S., Mackintosh, N. J. & Mackintosh, J. (1965). Shape and size discrimination in Octopus: the effects of pretraining along different dimensions. Journal of Genetic Psychology, 107, 110.Google Scholar
Sutherland, R. L., Mäthger, L. M., Hanlon, R. T., Urbas, A. M. & Stone, M. O. (2008a). Cephalopod coloration model. I. Squid chromatophores and iridophores. Journal of the Optical Society of America A, 25, 588599.CrossRefGoogle ScholarPubMed
Sutherland, R. L., Mäthger, L. M., Hanlon, R. T., Urbas, A. M. & Stone, M. O. (2008b). Cephalopod coloration model. II. Multiple layer skin effects. Journal of the Optical Society of America A, 25, 20442054.Google Scholar
Suzuki, M., Kimura, T., Ogawa, H., Hotta, K. & Oka, K. (2011). Chromatophore activity during natural pattern expression by the squid Sepioteuthis lessoniana: contributions of miniature oscillation. PLoS ONE, 6, e18244.Google Scholar
Sweeney, A. M., Haddock, S. H. D. & Johnsen, S. (2007). Comparative visual acuity of coleoid cephalopods. Integrative and Comparative Biology, 47, 808814.Google Scholar
Sweeney, M. J., Roper, C. F. E., Mangold, K. M., Clarke, M. R. & Boletzky, S. von (1992). Larval and juvenile cephalopods: a manual for their identification. Smithsonian Contributions to Zoology, 513, 1282.Google Scholar
Sykes, A. V., Pereira, D., Rodriguez, C., Lorenzo, A. & Andrade, J. P. (2013). Effects of increased tank bottom areas on cuttlefish (Sepia officinalis, L.) reproduction performance. Aquaculture Research, 44, 10171028.Google Scholar
Takami, T. & Suzu-Uchi, T. (1993). Southward migration of the Japanese common squid (Todarodes pacificus) from northern Japanese waters, in Recent Advances in Cephalopod Fisheries Biology (ed. Okutani, T., O’Dor, R. K. & Kubodera, T.), pp. 537543. Tokyo: Tokai University Press.Google Scholar
Taki, I. (1941). On keeping octopods in an aquarium for physiological experiments, with remarks on some operative techniques. Venus: Japanese Journal of Malacology, 10, 140156.Google Scholar
Talbot, C. M. & Marshall, J. (2010a). Polarization sensitivity in two species of cuttlefish – Sepia plangon (Gray 1849) and Sepia mestus (Gray 1849) – demonstrated with polarized optomotor stimuli. Journal of Experimental Biology, 213, 33643370.Google Scholar
Talbot, C. M. & Marshall, J. (2010b). Polarization sensitivity and retinal topography of the striped pyjama squid (Sepioloidea lineolata – Quoy/Gaimard 1832). Journal of Experimental Biology, 213, 33713377.Google Scholar
Talbot, C. M. & Marshall, J. N. (2011). The retinal topography of three species of coleoid cephalopod: significance for perception of polarized light. Philosophical Transactions of the Royal Society B – Biological Sciences, 366, 724733.Google Scholar
Tansley, K. (1965). Vision in Vertebrates. London: Chapman & Hall.Google Scholar
Tao, A. R., DeMartini, D. G., Izumi, M. et al. (2010). The role of protein assembly in dynamically tunable bio-optical tissues. Biomaterials, 31, 793801.Google Scholar
Tardent, P. (1962). Keeping Loligo vulgaris L. in the Naples aquarium. 1st Congres International d’Aquariologie, A, 4146.Google Scholar
Tasaki, K. & Karita, K. (1966). Intraretinal discrimination of horizontal and vertical planes of polarized light by Octopus. Nature, 209, 934.Google Scholar
Taylor, M. A. (1986). Stunning whales and deaf squids. Nature, 323, 298299.Google Scholar
Taylor, P. B. & Chen, L. C. (1969). The predator–prey relationship between the octopus (Octopus bimaculatus) and the California scorpionfish (Scorpaena guttata). Pacific Science, 23, 311316.Google Scholar
Teichert, C. (1988). Main features of cephalopod evolution, in The Mollusca, Vol. 12: Paleontology and Neontology of Cephalopods. (ed. Clarke, M. R. & Trueman, E. R.), pp. 1179. San Diego, CA: Academic Press.Google Scholar
Temple, S. E., Pignatelli, V., Cook, T. et al. (2012). High-resolution polarisation vision in a cuttlefish. Current Biology, 22, R121R122.Google Scholar
Terrace, H. S., Petitto, L. A., Sanders, R. J. & Bever, T. G. (1979). Can an ape create a sentence? Science, 200, 891902.Google Scholar
Thayer, G. H. (1909). Concealing-Coloration in the Animal Kingdom. An Exposition of the Laws of Disguise through Color and Pattern: Being a Summary of Abbott H. Thayer’s Discoveries. New York: The Macmillan Company.Google Scholar
Thomas, R. F. (1977). Systematics, distribution, and biology of cephalopods of the genus Tremoctopus (Octopoda: Tremoctopodidae). Bulletin of Marine Science, 27, 353392.Google Scholar
Thompson, J. T. & Voight, J. R. (2003). Erectile tissue in an invertebrate animal: the Octopus copulatory organ. Journal of Zoology, 261, 101108.Google Scholar
Thorpe, W. H. (1963). Learning and Instinct in Animals. London: Methuen.Google Scholar
Tinbergen, L. (1939). Zur Fortpflanzungsethologie von Sepia officinalis L. Archives Néerlandaises de Zoologie, 3, 323364.Google Scholar
Tinbergen, L. (1960). The natural control of insects in pinewoods. I. Factors influencing the intensity of predation by songbirds. Archives Néerlandaises Zoologie, 13, 265343.Google Scholar
Tinbergen, N. (1951). The Study of Instinct. Oxford: Clarendon Press.Google Scholar
Tinbergen, N. (1959). Comparative studies of the behaviour of gulls (Laridae): a progress report. Behaviour, 15, 170.Google Scholar
Tinbergen, N. (1963). On aims and methods of ethology. Zeitschrift für Tierpsychologie, 20, 410433.Google Scholar
Toll, R. B. (1983). The lycoteuthid genus Oregoniateuthis Voss, 1956, a synonym of Lycoteuthis Pfeffer, 1900 (Cephalopoda: Teuthoidea). Proceedings of the Biological Society of Washington, 96, 365369.Google Scholar
Toll, R. B. & Hess, S. C. (1981). Cephalopods in the diet of the swordfish, Xiphias gladius, from the Florida Straits. Fishery Bulletin, 79, 765774.Google Scholar
Tollrian, R. & Harvell, C. D. (ed.) (1999). The Ecology and Evolution of Inducible Defenses. Princeton, NJ: Princeton University Press.Google Scholar
Tong, D., Rozas, N. S., Oakley, T. H. et al. (2009). Evidence for light perception in a bioluminescent organ. Proceedings of the National Academy of Sciences USA, 106, 93869841.Google Scholar
Tranter, D. J. & Augustine, O. (1973). Observations on the life history of the blue-ringed octopus Hapalochlaena maculosa. Marine Biology, 18, 115128.Google Scholar
Tricarico, E., Borrelli, L., Gherardi, F. & Fiorito, G. (2011). I know my neighbour: individual recognition in Octopus vulgaris. PLoS ONE, 6.Google Scholar
Troscianko, T., Benton, C. P., Lovell, P. G., Tolhurst, D. J. & Pizlo, Z. (2009). Camouflage and visual perception. Philosophical Transactions of the Royal Society B – Biological Sciences, 364, 449461.Google Scholar
Trueman, E. R. & Packard, A. (1968). Motor performances of some cephalopods. Journal of Experimental Biology, 49, 495507.CrossRefGoogle Scholar
Tsuchiya, K. & Uzu, T. (1997). Sneaker male in octopus. Venus: Japanese Journal of Malacology, 56, 177181.Google Scholar
Tublitz, N. J., Gaston, M. R. & Loi, P. K. (2006). Neural regulation of a complex behavior: body patterning in cephalopod molluscs. Integrative and Comparative Biology, 46, 880889.Google Scholar
Tucker, J. K. & Mapes, R. H. (1978). Possible predation on Nautilus pompilius. Veliger, 21, 9598.Google Scholar
Tyrie, E. K., Hanlon, R. T., Siemann, L. A. & Uyarra, M. C. (2015). Coral reef flounders, Bothus lunatus, choose substrates on which they can achieve camouflage with their limited body pattern repertoire. Biological Journal of the Linnean Society, 114, 629638.Google Scholar
Uchiyama, K. & Tanabe, K. (1999). Hatching of Nautilus macromphalus in the Toba aquarium, Japan, in Advancing Research on Living and Fossil Cephalopods (ed. Olóriz, F. & Rodríguez-Tovar, F. J.), pp. 1116. New York: Kluwer Academic.Google Scholar
Ueda, K. (1985). Studies on the growth, maturation and migration of the shiriyake-ika, Sepiella japonica Sasaki. Bulletin of the Nansei Regional Fisheries Research Laboratory, 19, 142.Google Scholar
Uexküll, J. von (1905). Leitfaden in das Studium des Experimentale Biologie der Wassertiere. Wiesbaden: J.F. Bergmann.Google Scholar
Ulmer, K. M., Buresch, K. C., Kossodo, M. M. et al. (2013). Vertical visual features have a strong influence on cuttlefish camouflage. The Biological Bulletin, 224, 110118.Google Scholar
Vallin, A., Dimitrova, M., Kodandaramaiah, U. & Merilaita, S. (2011). Deflective effect and the effect of prey detectability on anti-predator function of eyespots. Behavioral Ecology and Sociobiology, 65, 16291636.Google Scholar
van Camp, L. M., Donnellan, S. C., Dyer, A. R. & Fairweather, P. G. (2004). Multiple paternity in field- and captive-laid egg strands of Sepioteuthis australis (Cephalopoda: Loliginidae). Marine and Freshwater Research, 55, 819823.Google Scholar
van Camp, L. M., Fairweather, P. G., Steer, M. A., Donnellan, S. C. & Havenhand, J. N. (2005). Linking male and female morphology to reproductive success in captive southern calamary (Sepioteuthis australis). Marine and Freshwater Research, 56, 933941.Google Scholar
Van Heukelem, W. F. (1966). Some aspects of the ecology and ethology of Octopus cyanea Gray. M.S. Thesis, Honolulu, HI: University of Hawaii.Google Scholar
Van Heukelem, W. F. (1977). Laboratory maintenance, breeding, rearing and biomedical research potential of the Yucatan octopus (Octopus maya). Laboratory Animal Science, 27, 852859.Google Scholar
Van Heukelem, W. F. (1983a). Octopus cyanea, in Cephalopod Life Cycles, Vol. 1. (ed. Boyle, P. R.), pp. 267276. London: Academic Press.Google Scholar
Van Heukelem, W. F. (1983b). Octopus maya, in Cephalopod Life Cycles, Vol. 1 (ed. Boyle, P. R.), pp. 311323. London: Academic Press.Google Scholar
van Staaden, M. J., Searcy, W. A. & Hanlon, R. T. (2011). Signaling aggression, in Advances in Genetics (ed. Huber, R., Bannasch, D. L. & Brennan, P.), pp. 2349. Elsevier, Inc.Google Scholar
Vecchione, M. (1987). Juvenile ecology, in Cephalopod Life Cycles, Vol. 2: Comparative Reviews (ed. Boyle, P. R.), pp. 6184. London: Academic Press.Google Scholar
Vecchione, M. (1988). In-situ observations on a large squid-spawning bed in the eastern Gulf of Mexico. Malacologia, 29, 135141.Google Scholar
Vecchione, M. & Roper, C. F. E. (1991). Cephalopods observed from submersibles in the western North Atlantic. Bulletin of Marine Science, 49, 433445.Google Scholar
Vecchione, M. & Young, R. E. (2006). The squid family Magnapinnidae (Mollusca: Cephalopoda) in the Atlantic Ocean, with a description of a new species. Proceedings of the Biological Society of Washington, 119, 365372.Google Scholar
Vecchione, M., Roper, C. F. E., Widder, E. A. & Frank, T. M. (2002). In situ observations on three species of large-finned deep-sea squids. Bulletin of Marine Science, 71, 893901.Google Scholar
Vecchione, M., Young, R. E., Guerra, A. et al. (2001). Worldwide observations of remarkable deep-sea squids. Science, 294, 25052506.Google Scholar
Vermeij, G. J. (1987). Evolution and Escalation: An Ecological History of Life. Princeton: Princeton University Press.Google Scholar
Verne, J. (1869). Twenty Thousand Leagues under the Sea. Paris.Google Scholar
Verrill, A. E. (1880–81). The cephalopods of the northeastern coast of America. Part II. The smaller cephalopods, including the ‘squids’ and the octopi, with other allied forms. Transactions of the Connecticut Academy of Sciences, 5, 259446.Google Scholar
Verwoerd, D. J. (1987). Observations on the food and status of the Cape Clawless Otter Aonyx capensis at Betty’s Bay, South Africa. South African Journal of Zoology, 22, 3339.Google Scholar
Villanueva, R. (1992). Deep-sea cephalopods of the north-western Mediterranean: indications of up-slope ontogenetic migration in two bathybenthic species. Journal of Zoology (London), 227, 267276.Google Scholar
Villanueva, R. (1994). Decapod crab zoeae as food for rearing cephalopod paralarvae. Aquaculture, 128, 143152.Google Scholar
Villanueva, R. & Guerra, A. (1991). Food and prey detection in two deep-sea cephalopods: Opisthoteuthis agassizi and O. vossi (Octopoda: Cirrata). Bulletin of Marine Science, 49, 288299.Google Scholar
Vincent, T. L. S., Scheel, D. & Hough, K. R. (1998). Some aspects of diet and foraging behavior of Octopus dofleini (Wülker, 1910) in its northernmost range. Marine Ecology, 19, 1329.Google Scholar
Vitti, J. J. (2013). Cephalopod cognition in an evolutionary context: implications for ethology. Biosemiotics, 6, 393401.Google Scholar
Voight, J. R. (1991a). Ligula length and courtship in Octopus digueti: a potential mechanism of mate choice. Evolution, 45, 17261730.Google Scholar
Voight, J. R. (1991b). Enlarged suckers as an indicator of male maturity in Octopus. Bulletin of Marine Science, 49, 98106.Google Scholar
Voight, J. R. (1992). Movement, injuries and growth of members of a natural population of the Pacific pygmy octopus, Octopus digueti. Journal of Zoology (London), 228, 247263.Google Scholar
Voight, J. R. (1996). The hectocotylus and other reproductive structures of Berryteuthis magister (Teuthoidea: Gonatidae). Veliger, 39, 117124.Google Scholar
Voight, J. R. (2000). A deep-sea octopus (Graneledone cf. boreopacifica) as a shell-crushing hydrothermal vent predator. Journal of Zoology, 252, 335341.Google Scholar
Voight, J. R. (2001). The relationship between sperm reservoir and spermatophore length in benthic octopuses (Cephalopoda: Octopodidae). Journal of the Marine Biological Association of the United Kingdom, 81, 983986.Google Scholar
Voight, J. R. (2005). Hydrothermal vent octopuses of Vulcanoctopus hydrothermalis, feed on bathypelagic amphipods of Halice hesmonectes. Journal of the Marine Biological Association of the United Kingdom, 85, 985988.Google Scholar
Voight, J. R. (2008). Observations of deep-sea octopodid behavior from undersea vehicles. American Malacological Bulletin, 24, 4350.Google Scholar
Voight, J. R. & Feldheim, K. A. 2009 Microsatellite inheritance and multiple paternity in the deep-sea octopus Graneledone boreopacifica (Mollusca: Cephalopoda). Invertebrate Biology 128, 2630.Google Scholar
Voight, J. R., Portner, H. O. & O’Dor, R. K. (1994). A review of ammonia-mediated buoyancy in squids (Cephalopoda: Teuthoidea). Marine and Freshwater Behaviour and Physiology, 25, 193203.Google Scholar
von Byern, J. & Klepal, W. (2006). Adhesive mechanisms in cephalopods: a review. Biofouling, 22, 329338.Google Scholar
von Byern, J., Scott, R., Griffiths, C. et al. (2011). Characterization of the adhesive areas in Sepia tuberculata (Mollusca, Cephalopoda). Journal of Morphology, 272, 12451258.Google Scholar
von Byern, J., Wani, R., Schwaha, T., Grunwald, I. & Cyran, N. (2012). Old and sticky – adhesive mechanisms in the living fossil Nautilus pompilius (Mollusca, Cephalopoda). Zoology, 115, 111.Google Scholar
Voss, G. L. (1956). A review of the cephalopods of the Gulf of Mexico. Bulletin of Marine Science, 6, 85178.Google Scholar
Voss, N. A. (1980). A generic revision of the Cranchiidae (Cephalopoda, Oegopsida). Bulletin of Marine Science, 30, 365412.Google Scholar
Vovk, A. N. (1974). Feeding habits of the North American squid Loligo pealei Lesueur. Fisheries and Marine Service, Translation Series, 3304, 114.Google Scholar
Vovk, A. N. (1977). The position of the longfin squid Loligo pealei Les. in the ecosystem. Fisheries and Marine Service, Translation Series, 3977, 113.Google Scholar
Vovk, A. N. & Khvichiya, L. A. (1980). On feeding of long-finned squid (Loligo pealei) juveniles in Subareas 5 and 6. North Atlantic Fisheries Organization Scientific Council Research Document, 80 /VI/50, N087, 19.Google Scholar
Wada, T., Takegaki, T., Mori, T. & Natsukari, Y. (2005a). Alternative male mating behaviors dependent on relative body size in captive oval squid Sepioteuthis lessoniana (Cephalopoda, Loliginidae). Zoological Science, 22, 645651.Google Scholar
Wada, T., Takegaki, T., Mori, T. & Natsukari, Y. (2005b). Sperm displacement behavior of the cuttlefish Sepia esculenta (Cephalopoda: Sepiidae). Journal of Ethology, 23, 8592.Google Scholar
Wada, T., Takegaki, T., Mori, T. & Natsukari, Y. (2006). Reproductive behavior of the Japanese spineless cuttlefish Sepiella japonica. Venus, 65, 221228.Google Scholar
Wada, T., Takegaki, T., Mori, T. & Natsukari, Y. (2010). Sperm removal, ejaculation and their behavioural interaction in male cuttlefish in response to female mating history. Animal Behaviour, 79, 613619.Google Scholar
Walderon, M. D., Nolt, K. J., Haas, R. E. et al. (2011). Distance chemoreception and the detection of conspecifics in Octopus bimaculoides. Journal of Molluscan Studies, 77, 309311.Google Scholar
Walker, J. J., Longo, N. & Bitterman, M. E. (1970). The octopus in the laboratory. Handling, maintenance, and training. Behavior Research Methods and Instrumentation, 2, 1518.Google Scholar
Waller, R. A. & Wicklund, R. I. (1968). Observations from a research submersible – mating and spawning of the squid, Doryteuthis plei. Bioscience, 18, 110111.Google Scholar
Walton, S. A., Korn, D. & Klug, C. (2010). Size distribution of the Late Devonian ammonoid Prolobites: indication for possible mass spawning events. Swiss Journal of Geosciences, 103, 475494.Google Scholar
Ward, A. J. W., Herbert-Read, J. E., Sumpter, D. J. T. & Krause, J. (2011). Fast and accurate decisions through collective vigilance in fish shoals. Proceedings of the National Academy of Sciences of the United States of America, 108, 23122315.Google Scholar
Ward, D. V. & Wainwright, S. A. (1972). Locomotory aspects of squid mantle structure. Journal of Zoology (London), 167, 437449.Google Scholar
Ward, P. D. (1987). The Natural History of Nautilus. London: Allen & Unwin.Google Scholar
Ward, P. D. & Bandel, K. (1987). Life history strategies in fossil cephalopods, in Cephalopod Life Cycles, Vol. 2: Comparative Reviews (ed. Boyle, P. R.), pp. 329350. London: Academic Press.Google Scholar
Ward, P. D. & Martin, A. W. (1980). Depth distribution of Nautilus pompilius in Fiji and Nautilus macromphalus in New Caledonia. The Veliger, 22, 259264.Google Scholar
Ward, P. D. & Saunders, W. B. (1997). Allonautilus: a new genus of living nautiloid cephalopod and its bearing on phylogeny of the Nautilida. Journal of Paleontology, 71, 10541064.Google Scholar
Ward, P. D. & Wicksten, M. K. (1980). Food sources and feeding behavior of Nautilus macromphalus. Veliger, 23, 119124.Google Scholar
Ward, P. D., Carlson, B., Weekly, M. & Brumbaugh, B. (1984). Remote telemetry of daily vertical and horizontal movement of Nautilus in Palau. Nature, 309, 248252.Google Scholar
Wardill, T. J., Gonzalez-Bellido, P. T., Crook, R. J. & Hanlon, R. T. (2012). Neural control of tuneable skin iridescence in squid. Proceedings of the Royal Society B – Biological Sciences, 279, 42434252.Google Scholar
Warnke, K. (1994). Some aspects of social interactions during feeding in Sepia officinalis (Mollusca) hatched and reared in the laboratory. Vie et Milieu, 44, 125131.Google Scholar
Warnke, K. M., Kaiser, R. & Hasselmann, M. (2012). First observations of a snail-like body pattern in juvenile Sepia bandensis (Cephalopoda: Sepiidae). A note. Neues Jahrbuch für Geologie und Palaontologie-Abhandlungen, 266, 5157.Google Scholar
Warnke, K. M., Meyer, A., Ebner, B. & Lieb, B. (2011). Assessing divergence time of Spirulida and Sepiida (Cephalopoda) based on hemocyanin sequences. Molecular Phylogenetics and Evolution, 58, 390394.Google Scholar
Warrant, E. (2004). Vision in the dimmest habitats on Earth. Journal of Comparative Physiology A – Neuroethology Sensory Neural and Behavioral Physiology, 190, 765789.Google Scholar
Warrant, E. J. (2007). Visual ecology: hiding in the dark. Current Biology, 17, R209R211.Google Scholar
Warren, L. R., Scheier, M. F. & Riley, D. A. (1974). Colour changes of Octopus rubescens during attacks on unconditioned and conditioned stimuli. Animal Behaviour, 22, 211219.Google Scholar
Wegener, B. J., Stuart-Fox, D. M., Norman, M. D. & Wong, B. B. M. (2013a). Strategic male mate choice minimizes ejaculate consumption. Behavioral Ecology, 24, 668671.Google Scholar
Wegener, B. J., Stuart-Fox, D., Norman, M. D. & Wong, B. B. M. (2013b). Spermatophore consumption in a cephalopod. Biology Letters, 9, DOI: 10.1098/rsbl.2013.0192.Google Scholar
Weihs, D. & Moser, H. G. (1981). Stalked eyes as an adaptation towards more efficient foraging in marine fish larvae. Bulletin of Marine Science, 31, 3136.Google Scholar
Wells, M. J. (1958). Factors affecting reactions to Mysis by newly hatched Sepia. Behaviour, 13, 96111.Google Scholar
Wells, M. J. (1959a). A touch learning centre in Octopus. Journal of Experimental Biology, 36, 590612.Google Scholar
Wells, M. J. (1959b). Functional evidence for neurone fields representing the individual arms within the central nervous system of Octopus. Journal of Experimental Biology, 36, 501511.Google Scholar
Wells, M. J. (1960). Proprioception and visual discrimination of orientation in Octopus. Journal of Experimental Biology, 37, 489499.Google Scholar
Wells, M. J. (1961). Weight discrimination by Octopus. Journal of Experimental Biology, 38, 127133.Google Scholar
Wells, M. J. (1962a). Brain and Behaviour in Cephalopods. London: Heinemann.Google Scholar
Wells, M. J. (1962b). Early learning in Sepia. Symposia of the Zoological Society of London, 8, 149169.Google Scholar
Wells, M. J. (1963). Taste by touch: some experiments with Octopus. Journal of Experimental Biology, 40, 187193.Google Scholar
Wells, M. J. (1964). Detour experiments with octopuses. Journal of Experimental Biology, 41, 621.Google Scholar
Wells, M. J. (1966). Learning in the Octopus. Symposia of the Society for Experimental Biology, 20, 477507.Google Scholar
Wells, M. J. (1967a). Sensitization and the evolution of associative learning, in Symposium on Neurobiology of Invertebrates, 1967, pp. 391411. Budapest: Hungarian Academy of Sciences.Google Scholar
Wells, M. J. (1967b). Short-term learning and interocular transfer in detour experiments with octopuses. Journal of Experimental Biology, 47, 393408.Google Scholar
Wells, M. J. (1978). Octopus: Physiology and Behaviour of an Advanced Invertebrate. London: Chapman and Hall.Google Scholar
Wells, M. J. (1990). Oxygen extraction and jet propulsion in cephalopods. Canadian Journal of Zoology, 68, 815824.Google Scholar
Wells, M. J. (1999). Why the ammonites snuffed it. Marine and Freshwater Behaviour and Physiology, 32, 103111.Google Scholar
Wells, M. J. & Clarke, A. (1996). Energetics: the costs of living and reproducing for an individual cephalopod. Philosophical Transactions of the Royal Society B: Biological Sciences, 351.Google Scholar
Wells, M. J. & O’Dor, R. K. (1991). Jet propulsion and the evolution of the cephalopods. Bulletin of Marine Science, 49, 419432.Google Scholar
Wells, M. J. & Wells, J. (1956). Tactile discrimination and the behaviour of blind Octopus. Pubblicazioni della Stazione Zoologica di Napoli, 28, 94126.Google Scholar
Wells, M. J. & Wells, J. (1957a). The function of the brain of Octopus in tactile discrimination. Journal of Experimental Biology, 34, 131142.Google Scholar
Wells, M. J. & Wells, J. (1957b). Repeated presentation experiments and the function of the vertical lobe in Octopus. Journal of Experimental Biology, 34, 469477.Google Scholar
Wells, M. J. & Wells, J. (1957c). The effect of lesions to the vertical and optic lobes on tactile discrimination in Octopus. Journal of Experimental Biology, 34, 378393.Google Scholar
Wells, M. J. & Wells, J. (1958). The effect of vertical lobe removal on the performance of octopuses in retention tests. Journal of Experimental Biology, 35, 337348.Google Scholar
Wells, M. J. & Wells, J. (1959). Hormonal control of sexual maturity in Octopus. Journal of Experimental Biology, 36, 133.Google Scholar
Wells, M. J. & Wells, J. (1970). Observations on the feeding, growth rate and habits of newly settled Octopus cyanea. Journal of Zoology (London), 161, 6574.Google Scholar
Wells, M. J. & Wells, J. (1972). Sexual displays and mating of Octopus vulgaris Cuvier and O. cyanea Gray and attempts to alter the performance by manipulating the glandular condition of the animals. Animal Behaviour, 20, 293308.CrossRefGoogle Scholar
Wells, M. J. & Wells, J. (1977). Cephalopoda: Octopoda, in Reproduction of Marine Invertebrates, Vol. 4: Molluscs: Gastropods and Cephalopods (ed. Giese, A. C. & Pearse, J. S.), pp. 291336. New York, NY: Academic Press.Google Scholar
Wells, M. J. & Young, J. Z. (1968). Learning with delayed rewards in Octopus. Zeitschrift für vergleichende Physiologie, 61, 103128.Google Scholar
Wells, M. J. & Young, J. Z. (1970). Stimulus generalisation in the tactile system of Octopus. Journal of Neurobiology, 2, 3146.Google Scholar
Wells, M. J., Freeman, N. H. & Ashburner, M. (1965). Some experiments on the chemotactile sense of octopuses. Journal of Experimental Biology, 43, 553563.Google Scholar
Wells, M. J., Wells, J. & O’Dor, R. K. (1992). Life at low oxygen tensions: the behaviour and physiology of Nautilus pompilius and the biology of extinct forms. Journal of the Marine Biological Association of the United Kingdom, 72, 313328.Google Scholar
Westermann, B. & Beuerlein, K. (2005). Y-maze experiments on the chemotactic behaviour of the tetrabranchiate cephalopod Nautilus pompilius (Mollusca). Marine Biology, 147, 145151.Google Scholar
Westermann, B. & Schipp, R. (1998). Cytological and enzyme-histochemical investigations on the digestive organs of Nautilus pompilius (Cephalopoda, Tetrabranchiata). Cell and Tissue Research, 293, 327336.Google Scholar
Westermann, B., Beck-Schildwachter, I., Beuerlein, K., Kaleta, E. F. & Schipp, R. (2004). Shell growth and chamber formation of aquarium-reared Nautilus pompilius (Mollusca, Cephalopoda) by X-ray analysis. Journal of Experimental Zoology, 301A, 930937.Google Scholar
Westermann, B., Ruth, P., Litzlbauer, H. D. et al. (2002). The digestive tract of Nautilus pompilius (Cephalopoda, Tetrabranchiata): an X-ray analytical and computational tomography study on the living animal. Journal of Experimental Biology, 205, 16171624.Google Scholar
Whitehead, M. R. (1990). Language and Literacy in the Early Years. London: Chapman.Google Scholar
Wickler, W. (1968). Mimicry. London: Weidenfeld and Nicolson.Google Scholar
Wickstead, J. (1956). An unusual method of capturing prey by a cuttlefish. Nature, 178, 929.Google Scholar
Wilbur, K. M. (1983–1988). Physiology of the Mollusca. San Diego: Academic Press.Google Scholar
Wiley, R. H. (1983). The evolution of communication: information and manipulation, in Communication (ed. Halliday, T. R. & Slater, P. J. B.), pp. 82113. Oxford: Blackwell.Google Scholar
Willey, A. (1899). General account of a zoological expedition to the South Seas during the years 1894–1897. Proceedings of the Zoological Society of London, 1899, 79.Google Scholar
Williams, B. L., Lovenburg, V., Huffard, C. L. & Caldwell, R. L. (2011). Chemical defense in pelagic octopus paralarvae: tetrodotoxin alone does not protect individual paralarvae of the greater blue-ringed octopus (Hapalochlaena lunulata) from common reef predators. Chemoecology, 21, 131141.Google Scholar
Williams, S. B., Pizarro, O., How, M. et al. (2009). Surveying nocturnal cuttlefish camouflage behaviour using an AUV. 2009 IEEE International Conference on Robotics and Automation, 214–219.Google Scholar
Williamson, G. R. (1965). Underwater observations of the squid Illex illecebrosus Lesueur in Newfoundland waters. Canadian Field-Naturalist, 79, 239247.Google Scholar
Williamson, R. (1988). Vibration sensitivity in the statocyst of the northern octopus, Eledone cirrosa. Journal of Experimental Biology, 134, 451454.Google Scholar
Williamson, R. (1991). Factors affecting the sensory response characteristics of the cephalopod statocyst and their relevance in predicting swimming performance. Biological Bulletin, 180, 221227.Google Scholar
Williamson, R. & Budelmann, B. U. (1991). Convergent inputs to octopus oculomotor neurones demonstrated in a brain slice preparation. Neuroscience Letters, 121, 215218.Google Scholar
Wilson, E. O. (1975). Sociobiology. Boston: Harvard University Press.Google Scholar
Wilson, M., Hanlon, R. T., Tyack, P. L. & Madsen, P. T. (2007). Intense ultrasonic clicks from echolocating toothed whales do not elicit anti-predator responses or debilitate the squid Loligo pealeii. Biology Letters, 3, 225227.Google Scholar
Winkelmann, I., Campos, P. F., Strugnell, J. et al. (2013). Mitochondrial genome diversity and population structure of the giant squid Architeuthis: genetics sheds new light on one of the most enigmatic marine species. Proceedings of the Royal Society B – Biological Sciences, 280.Google Scholar
Wirz, K. (1954). Études quantitatives sur le système nerveux des Céphalopodes. Comptes Rendus de L’Académie des Sciences, Paris, 238, 13531355.Google Scholar
Wirz, K. (1959). Étude biométrique du système nerveux des Céphalopodes. Bulletin Biologique de la France et de la Belgique, 93, 78117.Google Scholar
Wodinsky, J. (1969). Penetration of the shell and feeding on gastropods by Octopus. American Zoologist, 9, 9971010.Google Scholar
Wodinsky, J. (1973). Ventilation rate and copulation in Octopus vulgaris. Marine Biology, 20, 154164.Google Scholar
Wodinsky, J. (1978). Feeding behaviour of broody female Octopus vulgaris. Animal Behaviour, 26, 803813.Google Scholar
Wodinsky, J. (2008). Reversal and transfer of spermatophores by Octopus vulgaris and O. hummelinki. Marine Biology, 155, 91103.Google Scholar
Wolken, J. J. (1958). Retinal structure. Mollusc Cephalopods: Octopus, Sepia. J. Biophysics and Biochem. Cytol., 4.Google Scholar
Wood, F. G. (1963). Observations on the behavior of Octopus. Proceedings of the International Congress of Zoology, 16, 73.Google Scholar
Wood, J. B., Maynard, A. E., Lawlor, A. G. et al. (2010). Caribbean reef squid, Sepioteuthis sepioidea, use ink as a defense against predatory French grunts, Haemulon flavolineatum. Journal of Experimental Marine Biology and Ecology, 388, 2027.Google Scholar
Wood, J. B., Pennoyer, K. E. & Derby, C. D. (2008). Ink is a conspecific alarm cue in the Caribbean reef squid, Sepioteuthis sepioidea. Journal of Experimental Marine Biology and Ecology, 367, 1116.Google Scholar
Woodhams, P. L. & Messenger, J. B. (1974). A note on the ultrastructure of the octopus olfactory organ. Cell and Tissue Research, 152, 253258.Google Scholar
Woods, J. (1965). Octopus-watching off Capri. Animals, 7, 324327.Google Scholar
Wooton, R. J. (1990). Ecology of Teleost Fishes. London: Chapman and Hall.Google Scholar
Worms, J. (1983). World fisheries for cephalopods: a synoptic overview, in Advances in Assessment of World Cephalopod Resources (ed. Caddy, J. F.), pp. 120. Rome: FAO.Google Scholar
Wormuth, J. H. (1976). The Biogeography and Numerical Taxonomy of the Oegopsid Squid Family Ommastrephidae in the Pacific Ocean. Berkeley: University of California Press.Google Scholar
Würsig, B. (1986). Delphinid foraging strategies, in Dolphin Cognition and Behavior: A Comparative Approach (ed. Schusterman, R. J., Thomas, J. A. & Wood, F. G.), pp. 347359. Hillsdale: Lawrence Erlbaum Associates.Google Scholar
Xavier, J. C. & Cherel, Y. (2009). Cephalopod Beak Guide for the Southern Ocean. Cambridge: British Antarctic Survey.Google Scholar
Xavier, J. C., Cherel, Y., Assis, C. A., Sendao, J. & Borges, T. C. (2010). Feeding ecology of conger eels (Conger conger) in north-east Atlantic waters. Journal of the Marine Biological Association of the United Kingdom, 90, 493501.Google Scholar
Yamamoto, M. (1985). Ontogeny of the visual system in the cuttlefish, Sepiella japonica. 1. Morphological differentiation of the visual cell. Journal of Comparative Neurology, 232, 347361.Google Scholar
Yang, W. T., Hanlon, R. T., Krejci, M. E., Hixon, R. F. & Hulet, W. H. (1983). Laboratory rearing of Loligo opalescens, the market squid of California. Aquaculture, 31, 7788.Google Scholar
Yang, W. T., Hanlon, R. T., Lee, P. G. & Turk, P. E. (1989). Design and function of closed seawater systems for culturing loliginid squids. Aquacultural Engineering, 8, 4765.Google Scholar
Yang, W. T., Hixon, R. F., Turk, P. E. et al. (1986). Growth, behavior, and sexual maturation of the market squid, Loligo opalescens, cultured through the life cycle. Fishery Bulletin, 84, 771798.Google Scholar
Yano, K. & Tanaka, S. (1984). Some biological aspects of the deep sea squaloid shark Centroscymus from Suruga Bay, Japan. Bulletin of the Japanese Society of Scientific Fisheries, 50, 249.Google Scholar
Yarnall, J. L. (1969). Aspects of the behaviour of Octopus cyanea Gray. Animal Behaviour, 17, 747754.Google Scholar
Young, J. Z. (1939). Fused neurons and synaptic contacts in the giant nerve fibres of cephalopods. Philosophical Transactions of the Royal Society of London B, 229, 465503.Google Scholar
Young, J. Z. (1950). The Life of Vertebrates. Oxford: Oxford University Press.Google Scholar
Young, J. Z. (1958). Effect of removal of various amounts of the vertical lobes on visual discrimination by Octopus. Proceedings of the Royal Society B – Biological Sciences, 149, 441462.Google Scholar
Young, J. Z. (1960a). Observations on Argonauta and especially its method of feeding. Proceedings of the Zoological Society of London, 133, 471479.Google Scholar
Young, J. Z. (1960b). The statocysts of Octopus vulgaris. Proceedings of the Royal Society B: Biological Sciences, 152, 329.Google Scholar
Young, J. Z. (1960c). Unit processes in the formation of representations in the memory of Octopus. Proceedings of the Royal Society B: Biological Sciences, 153, 117.Google Scholar
Young, J. Z. (1960d). The failures of discrimination learning following the removal of the vertical lobes in Octopus. Proceedings of the Royal Society B: Biological Sciences, 153, 1846.Google Scholar
Young, J. Z. (1962a). Courtship and mating by a coral reef octopus (Octopus horridus). Proceedings of the Zoological Society of London, 138, 157162.Google Scholar
Young, J. Z. (1962b). Why do we have two brains?, in Interhemispheric Relations and Cerebral Dominance (ed. Mountcastle, V. B.), pp. 724. Baltimore: Johns Hopkins Press.Google Scholar
Young, J. Z. (1963). Light- and dark-adaptation in the eyes of some cephalopods. Proceedings of the Zoological Society of London, 140, 255272.Google Scholar
Young, J. Z. (1964a). A Model of the Brain. Oxford: Clarendon Press.Google Scholar
Young, J. Z. (1964b). Paired centres for the control of attack by Octopus. Proceedings of the Royal Society B: Biological Sciences, 159, 565588.Google Scholar
Young, J. Z. (1965a). The central nervous system of Nautilus. Philosophical Transactions of the Royal Society of London B, 249, 125.Google Scholar
Young, J. Z. (1965b). The organization of a memory system. Proceedings of the Royal Society B: Biological Sciences, 163, 285320.Google Scholar
Young, J. Z. (1966). The Memory System of the Brain. Oxford: Oxford University Press.Google Scholar
Young, J. Z. (1970). Stalked eyes of Bathothauma (Mollusca, Cephalopoda). Journal of Zoology, 162, 437447.Google Scholar
Young, J. Z. (1971). The Anatomy of the Nervous System of Octopus vulgaris. Oxford: Clarendon Press.Google Scholar
Young, J. Z. (1974). The central nervous system of Loligo. I. The optic lobe. Philosophical Transactions of the Royal Society of London B, 267, 263302.Google Scholar
Young, J. Z. (1976). The nervous system of Loligo. II. Suboesophageal centres. Philosophical Transactions of the Royal Society of London B, 274, 101167.Google Scholar
Young, J. Z. (1977a). Brain, behaviour and evolution of cephalopods. Symposia of the Zoological Society of London, 38, 377434.Google Scholar
Young, J. Z. (1977b). The nervous system of Loligo. III. Higher motor centres: the basal supraoesophageal lobes. Philosophical Transactions of the Royal Society of London B, 276, 351398.Google Scholar
Young, J. Z. (1978). Programs of the Brain. Oxford: Clarendon Press.Google Scholar
Young, J. Z. (1979). The nervous system of Loligo. V. The vertical lobe complex. Philosophical Transactions of the Royal Society of London B, 285, 311354.Google Scholar
Young, J. Z. (1981). The Life of Vertebrates. Oxford: Clarendon Press.Google Scholar
Young, J. Z. (1983). The distributed tactile memory system of Octopus. Proceedings of the Royal Society B: Biological Sciences, 218, 135176.Google Scholar
Young, J. Z. (1984). The statocysts of cranchiid squids (Cephalopoda). Journal of Zoology (London), 203, 121.Google Scholar
Young, J. Z. (1985). Cephalopods and neuroscience. Biological Bulletin, 168, 153158.Google Scholar
Young, J. Z. (1988). Evolution of the cephalopod brain, in The Mollusca, Vol.12: Paleontology and Neontology of Cephalopods (ed. Clarke, M. R. & Trueman, E. R.), pp. 215228. San Diego: Academic Press.Google Scholar
Young, J. Z. (1989). The angular acceleration receptor system of diverse cephalopods. Philosophical Transactions of the Royal Society of London B, 325, 189238.Google Scholar
Young, J. Z. (1991). Computation in the learning system of cephalopods. Biological Bulletin, 180, 200208.Google Scholar
Young, M., Kvitek, R. G., Iampietro, P. J., Hanlon, R. T. & Malliet, R. (2011). Seafloor mapping and landscape ecology analyses used to monitor variations in spawning site preference and benthic egg mop abundance for the California market squid (Doryteuthis opalescens). Journal of Experimental Marine Biology and Ecology, 407, 226233.Google Scholar
Young, R. E. (1972a). Brooding in a bathypelagic octopus. Pacific Science, 26, 400404.Google Scholar
Young, R. E. (1972b). The systematics and areal distribution of pelagic cephalopods from the seas off southern California. Smithsonian Contributions to Zoology, 97, 1159.Google Scholar
Young, R. E. (1973). Information feedback from photophores and ventral countershading in mid-water squid. Pacific Science, 27, 17.Google Scholar
Young, R. E. (1975a). Transitory eye shapes and the vertical distribution of two mid-water squids. Pacific Science, 29, 243255.Google Scholar
Young, R. E. (1975b). Function of the dimorphic eyes in the midwater squid Histioteuthis dofleini. Pacific Science, 29, 211218.Google Scholar
Young, R. E. (1977). Ventral bioluminescent countershading in midwater cephalopods. Symposia of the Zoological Society of London, 38, 161190.Google Scholar
Young, R. E. (1978). Vertical distribution and photosensitive vesicles of pelagic cephalopods from Hawaiian waters. Fishery Bulletin, 76, 583615.Google Scholar
Young, R. E. (1983). Oceanic bioluminescence: an overview of general functions. Bulletin of Marine Science, 33, 829845.Google Scholar
Young, R. E. (1995). Aspects of the natural history of pelagic cephalopods of the Hawaiian mesopelagic-boundary region. Pacific Science, 49, 143155.Google Scholar
Young, R. E. & Arnold, J. M. (1982). The functional morphology of a ventral photophore from the mesopelagic squid, Abralia trigonura. Malacologia, 23, 135163.Google Scholar
Young, R. E. & Bennett, T. M. (1988). Photophore structure and evolution within the Enoploteuthinae (Cephalopoda), in The Mollusca, Vol. 12: Paleontology and Neontology of Cephalopods (ed. Clarke, M. R. & Trueman, E. R.), pp. 241251. San Diego, CA.: Academic Press.Google Scholar
Young, R. E. & Harman, R. F. (1988). ‘Larva,’ ‘paralarva’ and ‘subadult’ in cephalopod terminology. Malacologia, 29, 201207.Google Scholar
Young, R. E. & Roper, C. F. E. (1976). Bioluminescent countershading in midwater animals: evidence from living squid. Science, 191, 10461048.Google Scholar
Young, R. E. & Roper, C. F. E. (1977). Intensity regulation of bioluminescence during countershading in living midwater animals. Fishery Bulletin, 75, 239252.Google Scholar
Young, R. E., Kampa, E. M., Maynard, S. D., Mencher, F. M. & Roper, C. F. E. (1980). Counterillumination and the upper depth limits of midwater animals. Deep-Sea Research Part A – Oceanographic Research Papers, 27, 671691.Google Scholar
Young, R. E., Roper, C. F. E. & Walters, J. F. (1979). Eyes and extraocular photoreceptors in midwater cephalopods and fishes: their roles in detecting downwelling light for counterillumination. Marine Biology, 51, 371380.Google Scholar
Young, R. E., Seapy, R. R., Mangold, K. & Hochberg, F. G. (1982). Luminescent flashing in the midwater squids Pterygioteithis microlampas and P. giardi. Marine Biology, 69, 299308.Google Scholar
Young, R. E., Vecchione, M. & Mangold, K. (2008). Cephalopoda Cuvier 1797: octopods, squids, cuttlefish, nautiluses. The Tree of Life Project, Version 21, April 2008.Google Scholar
Zahavi, A. (1975). Mate selection – a selection for handicap. Journal of Theoretical Biology, 53, 205214.Google Scholar
Zahavi, A. (1980). Ritualization and the evolution of movement signals. Behaviour, 72, 7781.Google Scholar
Zahavi, A. (1987). The theory of signal selection and some of its implications, in International Symposium on Biological Evolution (ed. Delfino, V. P.). Bari, Italy: Adriatica Editrice.Google Scholar
Zakharov, Y. D., Shigeta, Y., Smyshlyaeva, O. P., Popov, A. M. & Ignatiev, A. V. (2006). Relationship between delta C-13 and delta O-19 values of the recent Nautilus and brachiopod shells in the wild and the problem of reconstruction of fossil cephalopod habitat. Geosciences Journal, 10, 331345.Google Scholar
Zann, L. P. (1984). The rhythmic activity of Nautilus pompilius, with notes on its ecology and behavior in Fiji. Veliger, 27, 1928.Google Scholar
Zatylny, C., Gagnon, J., Boucaud-Camou, E. & Henry, J. (2000). ILME: A waterborne pheromonal peptide released by the eggs of Sepia officinalis. Biochemical and Biophysical Research Communications, 275, 217222.Google Scholar
Zeidberg, L. D. (2009). First observations of ‘sneaker mating’ in the California market squid, Doryteuthis opalescens (Cephalopoda: Myopsida). Marine Biodiversity Records, 2, e6.Google Scholar
Zeidberg, L. D., Butler, J. L., Ramon, D. et al. (2012). Estimation of spawning habitats of market squid (Doryteuthis opalescens) from field surveys of eggs off central and southern California. Marine Ecology – An Evolutionary Perspective, 33, 326336.Google Scholar
Zeidberg, L. D., Hamner, W., Moorehead, K. & Kristof, E. (2004). Egg masses of Loligo opalescens (Cephalopoda: Myopsida) in Monterey Bay, California following the El Nino event of 1997–1998. Bulletin of Marine Science, 74, 129141.Google Scholar
Zonana, H. V. (1961). Fine structure of squid retina. Bulletin of Johns Hopkins Hospital, 109, 185205.Google Scholar
Zullo, L., Sumbre, G., Agnisola, C., Flash, T. & Hochner, B. (2009). Nonsomatotopic organization of the higher motor centers in Octopus. Current Biology, 19, 16321636.Google Scholar
Zylinski, S. & Johnsen, S. (2011). Mesopelagic cephalopods switch between transparency and pigmentation to optimize camouflage in the deep. Current Biology, 21, 19371941.Google Scholar
Zylinski, S. & Osorio, D. (2011). What can camouflage tell us about non-human visual perception? A case study of multiple cue use in cuttlefish (Sepia officinalis), in Animal Camouflage: Mechanisms and Function (ed. Stevens, M. & Merilaita, S.), pp. 164185. Cambridge: Cambridge University Press.Google Scholar
Zylinski, S., Darmaillacq, A. S. & Shashar, N. (2012). Visual interpolation for contour completion by the European cuttlefish (Sepia officinalis) and its use in dynamic camouflage. Proceedings of the Royal Society B – Biological Sciences, 279, 23862390.Google Scholar
Zylinski, S., How, M. J., Osorio, D., Hanlon, R. T. & Marshall, N. J. (2011). To be seen or to hide: visual characteristics of body patterns for camouflage and communication in the Australian giant cuttlefish Sepia apama. American Naturalist, 177, 681690.Google Scholar
Zylinski, S., Osorio, D. & Shohet, A. (2009a). Cuttlefish camouflage: context-dependent body pattern use during motion. Proceedings of the Royal Society B: Biological Sciences, 276, 39633969.Google Scholar
Zylinski, S., Osorio, D. & Shohet, A. (2009b). Edge detection and texture classification by cuttlefish. Journal of Vision, 9, 110.Google Scholar
Zylinski, S., Osorio, D. & Shohet, A. J. (2009c). Perception of edges and visual texture in the camouflage of the common cuttlefish, Sepia officinalis. Philosophical Transactions of the Royal Society B: Biological Sciences, 364, 439448.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org 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.

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.

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.

Available formats
×